JR. L... A??? 2.; . .........2....... .5... {.11 1.721451; T303: 4 oh 5 . m hi»... . 9 fin. 2“ I .2 u :4 .- | . . .t :2 h:r..n...5.la I: I.~'I|“Ba4K \ ~a~ 453 .3;l..‘ ‘- 5 a. 4k 1 . .1 gt, I. 2.5. ii. . . 12...“: . . . , 3...»... 33...“..... figuwcfanfiw..m . . ., . . .. . I. .. x ; .e , ..... 1. 1.31%.... E gs... .- ft: .395: Y ‘3'" § \ 5"" 'V‘V'fl‘l‘lfi l )3! i t. .3919. 5.1.3: 3-. httitavtb .531)..- t. 035.39.. ...‘n}; x 3-. I... fut-:16“: ii {£315.13}. . t . .ia .muiiyraz n. 6...... u... .a. . .z .. a... .1. .9... .53.... .E. .9 2;... .. 5‘1?! 3...). 9.13."...- littl 2C9! 81.. I 7 [.1 .lttai. . .1... it l;..!1.!3¢i1.l. Off?! 254.15... 1L 0:. )Dltt.‘ all; . f . : .. . , , . . 1...... . . .....~.....:.., 5+... , . . 3. 1| ”I. . i... THE->13 1.30m LIBRARY l michigan 8:3: Univeram l -....-.. This is to certify that the dissertation entitled THERMAL AND PHOTOCHEMICAL REACTIONS OF FORMALDEHYDE ADSORBED ON Cu(lOO) presented by TODD ROBERT BRYDEN has been accepted towards fulfillment of the requirements for Ph . D 0 degree in Chemis t ry Eww QMajor professor Date _Augna_t_9_._ZQQL_ MSU is an Affirmative Action/Equal Opportunity Institution 0-12771 PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE NOVM 020322 6/01 cJCIRC/DateDuapes-ois THERMAL AND PHOTOCHEMICAL REACTIONS OF FORMALDEHYDE ADSORBED ON Cu(lOO) By TODD ROBERT BRYDEN A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2001 ABSTRACT THERMAL AND PHOTOCHEMICAL REACTIONS OF FORMALDEHYDE ADSORBED ON Cu(lOO) By TODD ROBERT BRYDEN Thin films of polymers are promising candidates for a wide variety of technological applications including corrosion protection, chemical sensing media and photonic materials. Polymeric materials are attractive due to the potential of combining high chemical, mechanical, and thermal stability with the ability for synthetically tailoring electronic or optical properties. Current methods for producing polymeric thin films (Langmuir-Blodgett (LB) and self-assembled monolayers(SAM)) have provided few details of the polymerization reaction mechanism and thermal decomposition of the film. Additionally, many monomer systems are amenable to neither LB and SAM techniques nor direct adsorption due to the high reactivity of many surfaces. The use of a co-adsorbate to block reaction on these surfaces has the potential for controlled thin film deposition through photon and/or electron initiated polymerization. Adsorption of formaldehyde (HZCO) on Cu(100) at 85 K formed an overlayer of poly(oxymethylene) (POM). For coverages up to 1.06 (:0.22)x10‘5 H2CO molecules/cm2 (9:0.69), formaldehyde polymerized to form a monolayer of disordered POM arranged with the chain directions parallel to the surface plane. Formaldehyde formed POM with differing chain lengths; the long (on) and short ([3) chain species. These depolymerized to give two features in temperature—programmed desorption at approximately 207 and 219 K, respectively. The presence of two POM species has not been observed previously in studies of HZCO adsorption on metal surfaces. The complex desorption kinetics observed for the CL and B POM species were successfully modeled using equations based on the ratio of the average number of monomers unzipped from the chain per initiation event to the length of the polymer chain. Vibrational features were observed for the short chain (B) species at ~ 290, ~ 1020 and ~ 1120 cm'1 that can be assigned to the v(Cu-O), v(C-O) and p(CH3) modes of oxygen and methoxy endgroups, respectively. Pre-adsorbed methanol increases the proportion of short chain POM species by increasing the probability of termination for the B species. Lower thermal stability for adsorbed POM was observed compared to bulk POM and is believed to be related to the stability of the surface bound oxygen endgroup. Saturating the Cu(100) surface with carbon monoxide (CO) inhibited the thermal polymerization of HZCO on Cu(100). Formaldehyde weakly adsorbed on CO/Cu(100), desorbing at 104 K, corresponding to a desorption energy of 18.2 (:08) kJ/mol. Irradiation of the H2CO/CO/Cu(100) surface caused the molecularly adsorbed HZCO to polymerize forming (POM). Irradiation also caused the formation of ethylene glycol. Vibrational features observed after irradiation at 870 and 3365 cm'1 were assigned to v(CC) and v(OH) modes, respectively, of ethylene glycol indicating that it was formed promptly upon irradiation. The presence of ethylene glycol was confirmed by studying its adsorption on clean and oxygen—covered Cu(100). The formation of ethylene glycol was likely governed by geometric constraints present with the formaldehyde overlayer. To my barber... iv ACKNOWLEDGEMENTS There are so many people that deserve my thanks during the completion of this thesis. Of course, none of the work described here could have been done without the help and guidance of my advisor, Simon Garrett. When starting this work, I knew almost nothing about surface science and you provided the perfect atmosphere for daily discovery by always having your office door open for questions and discussions. From you I’ve also learned that “try it and see” is an appropriate reason for an experiment and that jaunty angles must be avoided at all costs when designing chambers and gas manifolds. I must also thank the other members of my committee, Professors Blanchard, Baker and Swain, for listening to my ideas, answering my numerous questions and providing constructive criticism when needed. Inheriting a nearly empty laboratory provides the opportunity to thank all those involved in the construction and repair of equipment. Russ Geyer deserves an award for all the silver soldering he did for us (and leak-tight on the first try). The rest of the machine shop (Sam, Glenn and Tom) and electronics shop (Ron, Scott and David) were also instrumental in the completion of this work. They always seemed to comply with my “emergency” requests. Drs. Per Askeland and Kathy Severin deserve a special thanks for making my transition into graduate school go smoothly. As the only remaining Ledford students, they had a wealth of information about the department and its inhabitants. The members of the Garrett group (Lili, Mike, Heather and Jason) were always open to discussions running the gamut from music, MSU sports, food, and occasionally, even science. They were often quite lively and I really enjoyed them. I should also thank the other members of the Garrett group (John and Mike) for keeping me abreast of the latest in fantasy sports. Lastly, none of the work would have been possible without the help from my family. Mom, Dad, Sue, Bob, Ann and Glenn were all instrumental in my success even though you may not have understood what I was doing here. Barbara was a saint during our time in East Lansing. Between working two or more jobs and putting up with my frustration during tough experimental sessions, she deserves most of the credit for helping me with the completion of this thesis through her never ending support. I love you and thank you for helping me achieve this goal. vi TABLE OF CONTENTS Page LIST OF TABLES ix LIST OF FIGURES x INTRODUCTION 1 1.1 Polymer Thin Films 1 1.2 Topochemical Reactions 2 1.3 Surface Topochemical Reactions 3 1.4 Influence of Co—adsorbate 6 1.5 Proposed System 7 1.6 Reaction of Formaldehyde with Metal Surfaces 10 1.7 Outline of Thesis 12 1.8 References 14 Chapter 2. Experimental 20 2.1 References 29 Chapter 3. Adsorption and Polymerization of Formaldehyde on Cu(100) 31 3.1 Introduction 32 3.2 Results and Discussion 33 3.3 Conclusions 55 3.4 References 56 Chapter 4. Evidence for Two Chain Length Distributions in the Thermal Polymerization of Formaldehyde on Cu( 100) 60 4.1 Introduction 61 4.2 Results and Discussion 62 4.3 Conclusions 88 4.4 References 90 Chapter 5. Photochemistry of Formaldehyde Adsorbed on CO-saturated Cu(100) 93 5.1 Introduction 94 5.2 Results and Discussion 95 5.3 Conclusions ‘ 116 5.4 References 117 Chapter 6. Conclusions and Future Work 120 6.1 Thermal Reactions and Control of Stability 120 6.2 Photochemical Reactions 123 6.3 References 127 vii APPENDICES Appendix A. Mass Spectra 130 Appendix B. Calculation of Electron Impact Ionization Cross-Sections 135 Appendix C. Electron Energy Loss Spectrometer and Operation Voltages 140 viii Table 3.1 Table 3.2 Table 4.1 Table 4.2 Table 5.1 Table 5.2 Table B.1 Table C.1 LIST OF TABLES Assignments of the vibrational bands (in cm") observed by EELS for 0.5 L and 21 L exposures of HzCO on Cu(100) at 85 K. Also shown are IR data for solid poly(oxymethylene) (POM) and H2CO. Assignments of the vibrational bands (in cm'l) observed by EELS for 0.7 L exposure of trioxane on Cu(100) at 85 K. Also shown is IR data for solid trioxane. Results of the fits to equations (4.5) and (4.8) for the data shown in Figures 4.2 and 4.3. Assignments of the vibrational losses (in cm") observed by EELS for 1 ML coverages of HzCO and D2CO on Cu(100) at 85 K; also shown are IR data for solid poly(oxymethylene-hz) (POM-hz) and poly(oxymethylene-dz) (POM-d2). Assignments of the vibrational bands (in cm'l) observed for a 1.1 ML coverage of HZCO on CO-saturated Cu(100) after 0 and 15 minutes UV irradiation. Also shown are IR data for solid HZCO and POM along with EELS data for POM on Ag(111). Assignments of the vibrational bands (in cm") for the species observed following 15 minutes UV irradiation and annealing to 270 K of a 1.1 ML coverage of HzCO on CO-saturated Cu(100). Also shown are IR data for liquid (CH20H)2 along with EELS data for (CHZOH)2 on O/Ag(110) and (CHZOH)2 on O/Cu(100). Molecular orbital constants calculated at the Hartree-Fock level of theory with a 6-31G* basis set. These constants were used to evaluate the electron impact ionization cross-section. Operating voltages for the ELS3000 electron energy loss spectrometer. Page 39 41 71 75 102 109 137 141 Figure 1.1 Figure 1.2 Figure 1.3 Figure 1.4 Figure 2.1 Figure 2.2 Figure 2.3 Figure 3.1 Figure 3.2 LIST OF FIGURES Schematic representation of a topochernical reaction on a solid lattice. Crystal structure of orthorhombic POM (o-POM). Filled circles are methylene (CH2) units. Projection of the o-POM b-c plane onto Cu( 100). Schematic representation of substrate-mediated photochemistry where adsorption of photons, by the substrate, either creates a) subvacuum electrons (hvwork function). Sample mount. Oxygen coverage on Cu(100), resulting from the directed dosing of O; at 300 K, as determined by XPS. Ultraviolet/Visible spectrum of ~ 1 M NiSO4 (aq) solution used as a filter for the medium-pressure Hg arc lamp. Mass 29 thermal desorption spectra for increasing doses of HzCO on Cu(100). All exposures were made at 85 K. The inset shows the total integrated area for the features between 190 and 230 K as a function of exposure. Exposures are: a) 0.2 L, b) 0.3 L, c) 0.5 L, d) 0.6 L, e) 0.8 L, f) 1.1 L, g) 1.8 L, h) 2.2 L, i) 2.7 L. Thermal desorption spectra for masses 2 (Hf), 18 (H20+), 28 (CO+), 29 (HCO+), 30 (H2CO+), and 31 (H2'3CO+) for a 1.0 L exposure of HZCO on Cu(100). Each spectrum is a separate 1.0 L exposure. All exposures were made at 85 K. Spectra are offset for clarity. Page 13 21 25 28 34 37 Figure 3.3 Figure 3.4 Figure 3.5 Figure 3.6 Figure 3.7 Figure 3.8 Figure 4.1 EEL spectra for increasing exposures of HZCO on Cu(100) at 85 K. The sample was flashed to 300 K between exposures. EEL spectra of a) 0.5 L exposure of HZCO on Cu(100) and b) 0.7 L exposure of trioxane on Cu(100). Both spectra were acquired at 85 K. EEL spectra as a function of annealing temperature for a 6.3 L exposure of H2CO on Cu(100). The “as dosed” spectrum is shown in a) while b) and c) were annealed to the indicated temperature. All spectra were recorded at 85 K. The shoulder at ~ 200 cm'1 on the elastic peak is the lattice mode associated with the spectrum shown in a). This peak is absent in b) and c). XP spectra of a) C (ls) and b) O (ls) for increasing exposures of H2CO on Cu(100) at 85 K. The sample was flashed to 300 K between exposures. Exposures were 0, 0.1, 0.2, 0.4, 0.5, 0.7, 0.9, 1.2, 2.1, 5.4, L. Vertical lines are drawn at 288.8 eV and 533.8 eV for the C (Is) and O (ls) regions, respectively. Calculated coverage as a function of exposure for the adsorption of H2CO on Cu( 100) at 85 K based on XPS C(ls) peak areas. Equation of linear fit is y=0.898*x-0.082. Enlarged view of the polymer features from the TPD spectra shown in fig. 1. A constant background was subtracted from each spectrum. Exposures ranged from 0.1 L to 1.1 L. Mass 29 (HCO+) TPD spectra for (a) 1 ML H2CO on Cu(100), (b) 1 ML H2CO after annealing to 209 K and cooling to 85 K prior to TPD and (c) 1 ML HZCO annealed to 209 K and dosed with HzCO to re-saturate the surface. A constant linear background was subtracted from each spectrum. xi 38 40 46 47 49 63 Figure 4.2 Figure 4.3 Figure 4.4 Figure 4.5 Figure 4.6 Figure 4.7 Figure 4.8 Figure 4.9 Mass 29 (HCO+) TPD spectra for a 1 ML coverage of H2CO on Cu(100) exposed at 85 K as a function of heating rate (B). The heating rates were (a) 1.5 Ks", (b) 3.8 Ks", (c) 5.6 Ks’1 and (d) 8.7 Ks". The solid line (—) is the raw data, the dotted lines (---) are the fits to equations (4.5) and (4.8) and the dashed line (----) is the sum of the fits. Mass 29 (HCOI) TPD spectra for increasing coverages (0) of HzCO on Cu(100) exposed at 85 K. The coverages were (a) 0.1 ML, (b) 0.3 ML and (c) 0.6 ML. The solid line (—) is the raw data, the dotted lines (---) are the fits to equations (4.5) and (4.8) and the dashed line (----) is the sum of the fits. Simulated TPD spectra showing the effect of varying KIJDp. The following constant parameters were used for all spectra: E = 90 kJmol'l, v = 1x10”, and [3 = 5.6 Ks". Electron energy loss spectra of (a) 1 ML HzCO on Cu(100) exposed at 85 K and (b) 1 ML HzCO annealed to 209 K and dosed with HZCO to re-saturate the surface. Electron energy loss spectra of 1 ML HzCO on Cu( 100) exposed at 85 K and annealed to (a) 140 K, (b) 194 K, (c) 209 K and (d) 300 K. Each spectrum is a separate 1 ML H2CO dose. Electron energy loss spectra of (a) 1 ML HzCO and (b) 1 ML DzCO followed by a brief anneal to 140 K. Electron energy loss spectra of (a) 1 ML HzCO and (b) 1 ML DZCO followed by a brief anneal to 209 K. Mass 29 (HCOI) TPD spectra after the adsorption of (a) 0.2 L methanol (CH3OH) on the Cu(100) surface pre-covered with 0.6 ML HzCO and (b) 0.6 ML HZCO on the Cu(100) surface pre-covered with 0.2 L CH3OH. The “0 L CH3OH” spectrum is the same for both (a) and (b). xii 67 68 70 72 74 77 79 83 Figure 4.10 Figure 5.1 Figure 5.2 Figure 5.3 Figure 5.4 Figure 5.5 Electron energy loss spectra of (a) 0.2 L CH3OH, (b) 0.6 ML HZCO dosed onto the Cu(100) surface pre-covered with 0.2 L CH3OH and (c) the spectrum in (b) annealed to 209 K. Temperature-programmed desorption spectra for (a) m=30 (H2CO+) and (b) m=30 (H2CO+) and m=34 (D2COD+) for a 1.1 ML coverage of HzCO (DZCO) on CO-saturated Cu(100) as a function of UV irradiation time. D2CO was used to reduce coincident mass fragment interferences. Electron energy loss spectra for a 1.1 ML coverage of HzCO on CO-saturated Cu(100) for (a) 0 minutes irradiation at 85 K, (b) 15 minutes irradiation at 85 K, (c) 15 minutes irradiation followed by an anneal to 270 K and (d) 15 minutes irradiation followed by an anneal to 450 K. Each spectrum constituted a separate 1.1 ML coverage of HzCO. The data were acquired at 85 K. Temperature-programmed desorption spectra for m=31 (CH20H+) following exposures of 0.1 L, 0.2 L and 0.5 L of ethylene glycol (CHZOH)2 on clean Cu(100). The inset is the TPD spectrum obtained for a 0.5 L exposure on a Cu(100) pre- covered with 0.1 ML oxygen showing a new feature desorbing at 347 K. Electron energy loss spectra for (a) 0.1 L exposure of ethylene glycol (CH20H)2 on clean Cu(100) and (b) a 0.5 L exposure of (CH20H)2 on a Cu(100) pre-covered with 0.1 ML of oxygen followed by annealing to 270 K. Electron energy loss spectra for (a) 1.1 ML coverage of HzCO on CO—saturated Cu(100) irradiated for 15 minutes followed by an anneal to 270 K and (b) a 0.5 L exposure of (CH20H); on a Cu(100) pre-covered with 0.1 ML of oxygen followed by an anneal to 270 K. xiii 85 96 100 105 107 111 Figure 5.6 Figure 6.1 Figure 6.2 Figure A.1 Figure A.2 Figure A.3 Figure A.4 Figure 13.1 Electron energy loss spectra for (a) 1.1 ML coverage of DZCO on CO-saturated Cu(100) irradiated for 15 minutes followed by an anneal to 270 K and (b) a 0.5 L exposure of (CD20H); on a Cu(100) pre-covered with 0.1 ML of oxygen followed by an anneal to 270 K. Mass=29 (HCO+) TPD spectra showing the effect of X-ray irradiation on 1.1 ML HzCO on CO/Cu(100). Irradiation times were a) 0, b) 84 and c) 130 seconds. The X-ray wavelength was the Al K0t (hv=1486.6 eV) line and operated at 300 W (20 mA, 15 kV). Mass=30 (H2CO+) TPD spectra for the UV irradiation of 1.1 ML HzCO adsorbed on a surface that had been pre-covered with a saturation of poly(oxymethylene) (0.69 ML) for a) 0 minutes and b) 15 minutes UV irradiation. Mass spectra of a) background vacuum prior to introduction of HZCO and b) after a pressure of ~ 1x10'9 Torr of H2CO was achieved within the chamber. Mass spectra of a) background vacuum prior to introduction of DZCO and b) after a pressure of ~ 1x10'9 Torr of D2CO was achieved within the chamber. Mass spectra of a) background vacuum prior to introduction of trioxane (C3H603) and b) after a pressure of ~ 5X10'9 Torr of trioxane was achieved within the chamber. Mass spectra of a) background vacuum prior to introduction of ethylene glycol ((CHZOH)2) and b) after a pressure of ~ 1x10'9 Torr of ethylene glycol was achieved within the chamber. Electron impact ionization cross-section (051) as a function of incident electron energy for H2CO and H20 and a comparison to previous work done for water showing the small error introduced by using a smaller basis set. xiv 112 124 126 131 132 133 134 138 Figure C.l Schematic of the ELS3000 electron energy loss spectrometer. 142 XV INTRODUCTION 1.1 Polymer Thin Films The desire to create unique surfaces with specific chemical or physical properties has focussed attention on monolayers and multilayers of organic molecules or polymers. Through innovative synthetic design, polymeric materials can combine high chemical, mechanical and thermal stability with tailored adhesion, wettability, tribological, electronic or optical properties. Hence, macromolecular thin films are promising candidates for a wide range of technological applications including corrosion protectionlv2 and chemical sensing media.3 Ordered polymer films with crystalline character are especially important for technological applications since they generally possess superior optical and electronic properties compared to their amorphous counterparts. Device performance is strongly dependent on defects and grain boundaries. In particular, interest in photonic materialsflr5 field-effect transistorsé»7 and light-emitting diodes8 has focussed a great deal of effort on the development of techniques for producing crystalline, macromolecular thin films on solid surfaces. The polymer thin film may be deposited onto a solid surface either pre-formed, or as a monomer adlayer followed by subsequent in situ polymerization. Methods investigated to date include the deposition and polymerization of unsaturated Langmuir-Blodgett (LB) films9’10 and, more recently, self-assembled monolayers (SAMs).“'16 In these cases, the physical structure of the polymer film is known to be influenced by the atomic arrangement of the underlying solid substrate. For example, deposition of poly(oxymethylene) (POM), -(CH2-O)n-—, from solution onto cleaved alkali halide single-crystals generates extended polymer chains aligned with specific substrate crystal directions.17 The preferred directions appear to strongly correlate with minimal lattice mismatch between the macromolecular chain and surface unit cells. While polymerization reactions are believed to be important in some catalytic carbon-carbon coupling reactions,18 there are relatively few fundamental studies of small molecule polymerization reactions on surfaces. There is some precedent for thermal polymerization of HZCO to POM on several metal surfaces: O/Ag(110),19 Ni(110),20 Pt(111),21 Pd(111),22 O/Rh(111),23 O/Pd(lll),24 NiO(100)25 and Cu(110).26 Thermal polymerization of acetaldehyde on O/Ag(111) has also been observed.” These polymerization reactions, along with catalytic coupling reactions, are generally thought to be controlled by the relative rates of diffusion of the adsorbed monomer, initiator species and/or growing polymer chain.28 1.2 Topochemical Reactions In contrast to the reactions described above, where surface diffusion largely controls the rate of the reaction, solid-state, “diffusionless” polymerization reactions have been observed. These reactions are termed topochemical because the structure of the products can be directly related to the geometric arrangement of the reactants.29 Topochemical reactions are selective, rapid and steroespecific.3O If the product crystal structure can be predicted from the monomer structure, such crystalline monomer to crystalline polymer transformations are considered to be topotactic in nature.29 For both topochemical and topotactic reactions, a minimal amount of nuclear motion is observed and empirical rules have been developed from experimental and theoretical investigations: reactive centers must be S 4 A apart and unsaturated moieties must be nearly parallel.29r3' While the ideas of topochemical reactions may only be applicable in solid materials, it has been shown that a solid surface can exert a high degree of control over the polymerization and subsequent deposition of a polymer thin film. In such epitaxial polymerizations, structural information is conferred upon the first adsorbed polymer layer by the substrate and, in some cases, polymer morphologies are observed that are otherwise unobtainable from solution polymerization followed by crystallization.32‘35 1.3 Surface Topochemical Reactions We hope to use the substrate to order the monomer intentionally prior to polymerization. If the structure of the adsorbed monolayer matches closely with that of the polymer unit cell, the probability of a topochemical reaction occurring increases. We have chosen a system where a favorable packing geometry exists, as will be discussed below. Many small molecules are known to form well-ordered monolayers with high symmetry on transition metal surfaces increasing the possibility of parallel double bonds.18 Additionally, the nearest-neighbor distance of most of the transition metals is < 4 A,36 satisfying another of the rules for topochemical reactions. This is shown schematically in Figure 1.1 where the monomeric units are arranged on a lattice. Upon initiation, the monomer units are allowed free rotation and are oriented for reaction to form the polymer lattice. Initiation < —-> a K)/ M %/ gme XT\ monomer lattice polymer lattice Figure 1.1 Schematic representation of a topochemical reaction on a solid lattice. In terms of topochemical reactions, there are several differences between an adsorbed monolayer and a crystalline solid that will affect the polymerization reaction. Diffusion of adsorbates on metal surfaces is typically fast (at least 2 orders of magnitude greater than liquid diffusion coefficients) making the overlayer dynamic and less robust than the crystalline phase.37 The fast diffusion could disrupt the monomer order prior to polymerization, which will lessen the chance for topochemical reaction. Also, due to interadsorbate and adsorbate-substrate interactions, unique monolayer structures may be formed that may increase or decrease the probability of a topochemical reaction. For example, bulk crystalline trioxane has been observed to photopolymerize topochemically however the structure observed for monolayer trioxane adsorbed on Cu(l 1 1) (~ 15 A separation between molecules)38 would probably preclude such a reaction from occurring based on the empirical rules. Mechanistic factors will also affect the polymerization reaction on a surface. Successful creation of initiators and chain propagation along a specific direction in accordance with the structure of the overlayer will be crucial to formation of the polymer. The surface polymerization could suffer similar problems to solution polymerizations: inhibition of initiation and premature termination and/or chain transfer reactions. However, the rapid addition of monomers in a topotactic polymerization may be much faster than termination events. Indeed, in solid-state polymerization reactions, monomer addition may occur on a time scale as short as 10'13 5.39’40 Even in the topochemical polymerization of crystalline formaldehyde to crystalline poly(oxymethylene), where the monomer moves a large distance (~ 0.5 A), the addition of monomer was found to occur every 10'5 s at 80 K."'0 This is faster than the rate seen for solution-phase free radical polymerization where addition occurs on the time scale of 102-104 5 at room temperature.“ Using the surface to align reactants was first shown for ZHZS —-> H2+2HS42 and 2HX —> H2+X2 (X=Cl and Br)43 on LiF(001) where photochemical product energies and yields were markedly different than those seen in the gas-phase. The photochemistry of HX is thought to occur through an aligned HX dimer where the close proximity of the two HX molecules is afforded by adsorption on the LiF surface. Also, the photooxidation of CO on Pt(779) occurs by dissociation of co-adsorbed O; at step-edges, which are aligned to oxidize CO adsorbed at step-edges preferentially.44 The chemistry need not be photon driven. The observed probability of non-terminal addition of deuterium atoms to l-butene adsorbed on Cu(100) was found to be greater (~ 5x more) than that seen in the gas-phase and was ascribed to a steric effect based on the adsorption geometry of the molecule.45 For most of the systems which have been investigated to date, little is known about the adsorbed monolayer structure prior to polymerization, the polymer morphology or the details of the reaction mechanism. Moreover, the correlation between the atomic arrangement of the substrate and polymer has not been studied systematically. We are interested in elucidating the polymerization initiation, propagation and termination mechanisms operative for ordered monolayers of unsaturated small molecules, and the influence of the surface on the polymer film order, chain length, conformation and stability. Ultimately, control of some or all of these processes may enable the design of high-quality crystalline polymer thin films. To our knowledge, the possibility of controlling polymerization through the use of a surface acting as a reaction "template" (topotactic reaction) has not been investigated. 1.4 Influence of Co-adsorbate In addition to studying polymerization reactions on clean surfaces, co-adsorbed species offer the potential to control both the chain length and endgroups of the resulting polymer. In solution polymerizations, chain transfer agents are used to control the molecular weight distribution of a polymerization and determine the endgroup.41 On a surface, the co-adsorbate may participate in the reaction through initiation and/or termination events; in either case, the endgroup identity may be determined by the co- adsorbate. Endgroup stability of a polymer adsorbed on a surface is known to have a large impact on the thermal stability of the polymer. For example, bulk polyimide made from pyromellitic dianhydride (PMDA) and oxydianiline (ODA) is stable at temperatures up to 500 °C.""5v47 However, polyimide films made from reactive adsorption of PMDA and ODA on Cu(110) decompose at ~ 280 °C.48949 The thermal stability of the film is limited by the reactivity of the carboxylate endgroups bound to the surface, but the decomposition pathway and the influence of the substrate on the endgroup stability have not been correlated. A co—adsorbate may also be used to control reactivity of the substrate with respect to the monomer. Direct adsorption of the monomer is a simple alternative to LB and SAM methods. However some metal substrates are too reactive to form the polymer thin film through direct adsorption. For example, formation of POM films from H2CO is not possible through LB and SAM methods and direct adsorption of HzCO on Fe(100) causes decomposition, and not polymerization, of the adsorbed H2CO.50 Carbon monoxide is an obvious choice for use as the co-adsorbate to control the reactivity of the substrate because it forms a well-ordered monolayer on Cu(100) at 85 K51 and is only weakly chemisorbed, desorbing molecularly at ~ 180 K52 Co-adsorbed carbon monoxide (CO) is known to influence the chemistry observed on surfaces. For example, co-adsorbed CO has been found to increase the stability of ethylidyne on Ru(001),53 decrease the stability of the methyl hydrogen atoms of toluene on Ru(001),54 perturb the decomposition pathway of methylamine on Ru(001)55 and promote the decomposition of saturated hydrocarbons on Ni(755).56 1.5 Proposed System The system chosen for study is the adsorption of formaldehyde on Cu(100). Formaldehyde is an ideal molecule for studying polymerization reactions due to its low activation energy for polymerization (~ 12 kJ/mol) and the fact the reaction can be self- initiated without the use of an initiator species. Also, the gas, solution and solid-state reactions of formaldehyde have been studied extensively.39r40,57r53 Gaseous H2CO has a S1 (— 80 (1t* (- n) band origin at 3.49 eV57 and can dissociate, upon UV irradiation, via two channels: H2CO -—) H2 + CO (molecular) and H2CO —-> H' + HCO’ (radical). The molecular channel has its threshold at 3.52 eV. Polymerization is not expected to be initiated by the molecular photoproducts. The radical channel, whose threshold is at 3.73 eV, may initiate the polymerization through reaction of the radical species with molecular H2CO. Formaldehyde can be polymerized to form POM via cationic, anionic and free radical routes.58 Poly(oxymethylene) can also be formed through ring opening of the cyclic trimer (l,3,5-trioxane) and tetramer ( 1,3,5,7-tetraoxane) of H2CO. Both orthorhombic (o-POM) and trigonal (t-POM) crystal structures have been observed for POM in which the POM chains form helices.59 In t-POM, the helical conformation is designated D(1070’9) while the o-POM conformation is D(71’).60 The crystal structure of o- POM is shown in Figure 1.2. For t-POM, only one chain per unit cell is present. b=765A \ \ a=4J7A 38 fl M818 8% Figure 1.2 Crystal structure of orthorhombic POM (o-POM). Filled circles are methylene (CH2) units. c=356A The unit cell parameters of the Cu(100) surface match well with those of crystalline o-POM. The c dimension of the polymer unit cell (3.56 A) is only about 1 % smaller than the bulk copper unit cell (ao=3.61 A),36 and the b dimension (7.65 A) is about 6 % larger than 2a“. This is shown schematically in Figure 1.3 where the b-c plane is projected onto the Cu(100) surface plane. The close match of unit cells increases the probability for a topocherrrical reaction. The helical conformation observed in the solid-state may not be observed on the Cu(100) surface due to the strong adsorbate-surface interactions. The nearest-neighbor distance for copper (2.55 A) matches closely with that of the oxygen—oxygen repeat distance (2.49 A) for POM in a planar zigzag conformation. This conformation has not been observed experimentally due to dipole-dipole repulsion of the neighboring C-O-C units. The helical conformation has been calculated to be ~ 7 kJ/mol lower in energy than the planar zigzag structure.“ Figure 1.3 Projection of the o-POM b-c plane onto Cu(100). It has been shown that in the solid-state photochemistry of formaldehyde, a high degree of order prior to irradiation is essential for long chain growth of POM. Formaldehyde has been observed to undergo photopolymerization ('y-rays) in the solid- state down to temperatures as low as ~ 4 K producing high molecular weight, crystalline polymer.39r40. In contrast, no POM was detected after irradiation of amorphous films of H2CO at 10 and 77 K and, only after doping with C12 were oligomers formed (~ 6 monomer units)!“63 1.6 Reaction of Formaldehyde with Metal Surfaces As was mentioned previously, the adsorption of H2CO on metal surfaces has been investigated. Non-dissociative adsorption has been observed on clean Ag(110),19r64 Ag(111),65 Au(110)66 and Zn(0001)67 at temperatures below 100 K. Reactive adsorption of formaldehyde was found to occur on O/Ag(110),19v64 Ni(110),20 O/Zn(0001),67 Ru(001),68 Rh(111),23 Fe(100),50 Pt(l l 1),21 and Pd(111)22 with H2, CO and POM being formed. When POM has been observed, it has been speculated the interaction of H2CO with surface-bound oxygen forms either a forrnyl species(HCO)68 or a dioxymethylene (H2CO2)64 species which then participates in the polymerization process. The oxygen probably arises due to dissociation of the H2CO upon adsorption. Oxygen has the ability to act as a nucleophile as well as create Lewis acid sites in its immediate vicinity.64r69 Both can initiate the polymerization of formaldehyde.58 Also, partial thermal polymerization was observed on Cu(110) at 90 K but the effect of adsorbed oxygen was not addressed.26 10 There has been one investigation on the photopolymerization of formaldehyde adsorbed on a metal surface. Dai et al. polymerized formaldehyde adsorbed on Ag(111) using nanosecond pulses of 355 nm radiation.65r70’7l Formation of POM was observed at 80 K for coverages from 0.1 to l monolayer (ML) of adsorbed H2CO. In the absence of irradiation, no polymerization was observed. Also, no polymerization was observed upon exposure to 532 nm radiation suggesting a non-thermal initiation mechanism.70 Below 60 K, no polymerization occurred after irradiation, presumably due to reduced diffusion of H2CO or initiator species. However, polymerization did occur after these irradiated samples were annealed to higher temperatures?1 This is indicative of a diffusion-limited process and not a topochemical reaction. Recent results by Dai et al. have shown laser-induced polymerization of formaldehyde on Ag(1 11) occurs through photoexcited electrons generated within the substrate.72’73 Substrate sub-vacuum electrons have been observed to initiate chemistry of adsorbates as well as stimulate desorption.74‘77 This is represented schematically in Figure 1.4a where photons (hv — LUMO(1t*) hv Er: — HOMO (11) Metal HZCO a) Creation of subvacuum electrons followed by tunneling to LUMO of HZCO. C—V — LUMO (1t*) A vac — hv — HOMO (n) EF Metal H2CO b) Creation of vacuum electrons followed by attachment to HCO. Figure 1.4 Schematic representation of substrate—mediated photochemistry where adsorption of photons, by the substrate, either creates a) subvacuum electrons (hvwork function). observed to desorb from the surface and is believed to arise from H2CO dissociating on the clean Cu(100) substrate. A clean surface is obtained after heating to 300 K. Two different POM species are formed from the thermal polymerization of H2CO on Cu(100) at 85 K.79 This is the subject of Chapter 4. Two separate chain lengths are produced in the polymerization and the depolymerization kinetics have been modeled using standard depolymerization kinetic equations. The differences between the ratio of degree of polymerization and the average number of monomer units depolymerized at each initiation step separate the long and short chain POM species. Also, we have used co-adsorbed methanol to preferentially form the shorter species and confirm the identity of the endgroups. Chapter 5 describes the use of a co-adsorbate, carbon monoxide, to control the reactivity of the Cu(100) with respect to formaldehyde adsorption.80 A saturation coverage of CO inhibits the thermal polymerization of H2CO on Cu(100) at 85 K. Upon irradiation with UV photons, molecularly adsorbed H2CO reacts to form POM and ethylene glycol, which has not been observed previously in the photochemistry of adsorbed formaldehyde. The CO desorbs prior to POM depolymerization offering the possibility for forming patterned polymer films through photopolymerization, followed by desorption of monomer (from the unirradiated areas) and the CO spacer layer. 1.8 References (1) Laibinis, P.; Whitesides, G. J. Am. Chem. Soc. 1992, 114, 9022. (2) Zhao, M.; Liu, Y.; Crooks, R.; Bergbreiter, D. J. Am. Chem. Soc. 1999, 121, 923. 14 (3) Harsanyi, G. Polymer Films in Sensor Applications; TECHNOMIC: Lancaster, PA, 1995. (4) Lemoine, V.; Pocholle, J .; Le Barny, P.; Robin, P. In Molecular Nonlinear Optics; Zyss, 1., Ed.; Academic Press: Boston, 1994; pp 379. (5) Lytel, R.; Lipscomb, G.; Thackara, J.; Altman, J.; Elizondo, P.; Stiller, M.; Sullivan, B. In Nonlinear Optical and Electroactive Polymers; Prasad, P., Ullrich, D., Eds.; Plenum Press: New York, 1988; pp 415. (6) Katz, H.; Bao, Z. J. Phys. Chem. B. 2000, 104, 671. (7) Xu, 6.; Bao, Z.; Groves, J. Langmuir 2000, 16, 1834. (8) Grell, M.; Bradley, D. Adv. Mater. 1999, 11, 895. (9) Ejaz, M.; Yamamoto, S.; Ohno, K.; Tsujii, Y.; Fukuda, T. Macromolecules 1998, 31, 5934. (10) Shirai, E.; Urai, Y.; Itoh, K. J. Phys. Chem. B. 1998, 19, 3765. (l 1) Ford, J. F.; Vickers, T.; Mann, G; Schlenoff, J. Langmuir 1996, 12, 1944. (12) Sun, F.; Castner, D.; Grainger, D. Langmuir 1993, 9, 3200. (13) Kim, T.; Chan, K.; Crooks, R. J. Am. Chem. Soc. 1997, 119, 189. (14) Chan, K.; Kim, T.; Schoer, 1.; Crooks, R. J. Am. Chem. Soc. 1995, 117, 5875. (15) Balasubrarnanian, K.; Cammarata, V. Langmuir 1996, 12, 2035. (16) Batchelder, D.; Evans, 8.; Freeman, T.; Haussling, L.; Ringsdorf, H.; Wolf, H. J. Am. Chem. Soc. 1994, 116, 1050. (17) Mauritz, K.; Baer, E.; Hopfinger, A. J. Polym. Sci. : Macromol. Rev. 1978, 13, l. (18) Somorjai, G. Introduction to Surface Chemistry and Catalysis; Wiley and Sons: New York, 1994. (19) Stuve, E.; Madix, R.; Sexton, B. Surf. Sci. 1982, 119, 279. 15 (20) Richter, L.; Ho, W. J. Chem. Phys. 1985, 83, 2165. (21) Henderson, M.; Mitchell, G.; White, J. Surf. Sci. 1987, 188, 206. (22) Davis, J.; Barteau, M. J. Am. Chem. Soc. 1989, 111, 1782. (23) Houtrnan, C.; Barteau, M. Surf. Sci. 1991, 248, 57. (24) Davis, J.; Barteau, M. Surf. Sci. 1992, 268, 11. (25) Truong, C.; Wu, M.; Goodman, D. W. J. Am. Chem. Soc. 1993, 115, 3647. (26) Sexton, B.; Hughes, A.; Avery, N. Surf. Sci. 1985, I55, 366. (27) Sim, W.; Gardner, P.; King, D. J. Am. Chem. Soc. 1996, 118, 9953. (28) Bent, B. E. Chem. Rev. 1996, 96, 1361. (29) Thakur, M. In Encyclopedia of Polymer Science and Engineering; Wiley and Sons: New York, 1989; Vol. 15; pp 362. (30) Wright, J. Molecular Crystals; Cambridge University Press: Cambridge, 1987. (31) Salcedo, R.; Sansores, L.; Valladares, A.; Likhatchev, D.; Alecandrova, L.; Ogawa, T. Polymer 1996, 37, 1703. (32) Yamashita, Y.; Shimamura, K.; Kasahara, H.; Monobe, K. Syn. Met. 1987, 17, 253. (33) Lando, J.; Baer, E.; Rickert, S.; Nae, H.; Ching, S. In Initiation of Polymerization; 212 ed.; Bailey, F. E., Ed.; American Chemical Society: Washington, DC, 1983; Vol. 212. (34) Sano, M.; Lvov, Y.; Kunitaki, T. In Annual Review of Materials Science; Tirrell, M., Kaufman, E., Giordmaine, J., Wachtman, J., Eds.; Annual Reviews: Palo Alto, 1996; Vol. 26; PP 153. (35) Sano, M.; Sasaki, D.; Kunitake, T. Macromolecules 1992, 25, 6961. (36) CRC Handbook of Chemistry and Physics, 76 ed.; CRC Press: Boca Raton, 1995- 1996. (37) Seebauer, E.; Allen, C. Prog. Surf. Sci. 1995, 49, 265. (38) Hofmann, M.; Wegner, H.; Glenz, A.; Woll, C.; Grunze, M. J. Vac. Sci. Technol. A 1994, 12, 2063. (39) Gol’danskii, V.; Benderskii, V.; Trakhtenberg, L. In Advances in Chemical Physics; Progogine, 1., Rice, S., Eds.; Wiley and Sons: New York, 1989; Vol. 75; pp 349. (40) Gol’danskii, V. Ann. Rev. Phys. Chem. 1976, 27, 85. (41) Odian, G. Principles of Polymerization, 3 ed.; Wiley and Sons: New York, 1991. (42) Bourdon, E.; Das, P.; Harrison, 1.; Polanyi, J.; Segner, J.; Stanners, C.; Williams, R.; Young, P. Faraday Discuss. Chem. Soc. 1986, 82, 343. (43) Cho, C.; Polanyi, J.; Stanners, C. J. Chem. Phys. 1989, 90, 598. (44) Tripa, C.; Yates, J. J. Chem. Phys. 2000, 112, 2463. (45) Yang, M.; Teplyakov, A.; Bent, B. J. Phys. Chem. B. 1998, 102, 2985. (46) Karnbe, H. In Aspects of Degradation and Stabilization of Polymers; Jellinek, H., Ed.; Elsevier: Amsterdam, 1978; pp 393. (47) Feger, C.; Franke, H. In Polyimides: Fundamentals and Applications; Ghosh, M., Mittal, K., Eds.; Marcel Dekker: New York, 1996; pp 759. (48) Haq, S.; Richardson, N. J. Phys. Chem. B. 1999, 103, 5256. (49) Plank, R.; DiNardo, J .; Vohs, J. Phys. Rev. B. 1997, 55, 10241. (50) Hung, W.; Bemasek, 8. Surf. Sci. 1996, 346, 165-188. (51) Tracy, J. J. Chem. Phys. 1972, 56, 2748. (52) Bryden, T.; Garrett, S. J. Phys. Chem. B. 1999, 103, 10481. (53) Sasaki, T.; Kawada, F.; Aruga, T.; Iwasawa, Y. Surf. Sci. 1992, 278, 291. 17 (54) Rauscher, H.; Menzel, D. Surf. Sci. 1995, 342, 155. (55) Sasaki, T.; Aruga, T.; Kuroda, H.; Iwasawa, Y. Surf. Sci. 1992, 276, 69. (56) Orita, H.; Kondoh, H.; Nozoye, H. Chem. Phys. Lett. 1994, 228, 385. (57) Moore, C.; Weisshar, J. Ann. Rev. Phys. Chem. 1983, 34, 525. (58) Walker, J. F. Formaldehyde, 3rd. ed.; Reinhold Publishing Corp.: New York, 1964. (59) Tadokoro, H. In Macromolecular Reviews; Peterlin, A., Goodman, M., Okamura, S., Zimm, B., Mark, H., Eds.; Interscience: New York, 1967; Vol. 1; pp 119. (60) Helical symmetry is specified by the nomenclature X(2m1r/n) where X is the point group (C or D), m is the number of turns of the chain per unit cell and n is the number of monomer units per chain per unit cell). (61) Tadokoro, H.; Kobayashi, M.; Mori, K. Rep. Progr. Polymer Phys. Japan 1965, 8, 45. (62) Mansueto, E.; Ju, C.; Wight, C. J. Phys. Chem. 1989, 93, 2143-2147. (63) Manseuto, E.; Wight, C. J. Photochem. Photobiol. A:Chem. 1991, 60, 251. (64) Barteau, M.; Bowker, M.; Madix, B. Surf. Sci. 1980, 94, 303-322. (65) Fleck, L.; Ying, Z.; Feehery, M.; Dai, H—L. Surf. Sci. 1993, 296, 400. (66) Outka, D.; Madix, R. Surf. Sci. 1987, I 79, 361. (67) Sen, P.; Rao, C. Surf. Sci. 1986, 172, 269-280. (68) Anton, A.; Parrneter, J .; Weinberg, W. J. Am. Chem. Soc. 1986, 108, 1823. (69) Madix, R. Science 1986, 233, 1159-1166. (70) Fleck, L.; Feehery, W.; Plummer, E.; Ying, Z.; Dai, H. J. Phys. Chem. 1991, 95, 8428. 18 (71) Fleck, L.; Ying, Z.; Dai, H.-L. J. Vac. Sci. Technol. A 1993, 11, 1942. (72) Fleck, L.; Howe, P.; Kim, J .; Dai, H. J. Phys. Chem. 1996, 100, 8011. (73) Ying, Z.; Fleck, L.; Dai, H. In Laser Spectroscopy and Photochemistry on Metal Surfaces; Dai, H., Ho, W., Eds.; World Scientific Publishing Company: Singapore, 1995. (74) Mieher, W.; Ho, W. J. Chem. Phys. 1989, 91, 2755. (75) Zhou, X.; Coon, S.; White, J. J. Chem. Phys. 1990, 92, 1498. (76) Ying, 2.; Ho, W. Phys. Rev. Lett. 1990, 65, 741. (77) Hoffman, A.; Guo, X.; Yates, J.; Gadzuk, J.; Clark, C. J. Chem. Phys. 1989, 90, 5793. (78) Fleck, L.; Kim, J.; Dai, H.-L. Surf. Sci. 1996, 356, L417. (79) Bryden, T.; Garrett, S. Langmuir , in press. (80) Bryden, T.; Garrett, S. J. Phys. Chem. B. , in press. 19 Chapter 2 Experimental Ultrahigh Vacuum Chamber (UHV). The experiments to be described were conducted in two connected stainless steel ultrahigh vacuum chambers. Each chamber was separately pumped by a 270 Us ion pump. The first chamber, spherical in shape, was equipped with a 1-200 amu quadrupole mass spectrometer (QMS) (VG Masstorr 200 DX), an ion gun for noble gas ion sputtering (Phi 04-161 gun, 20-005F controller) and a molecular leak valve (Varian 9515106). This chamber was also equipped with a dual anode X-ray source (VG 3EXR2) and hemispherical electron energy analyzer (VG CLAM2) for X-ray photoelectron spectroscopy (XPS). The pressure was measured by a nude ion gauge. After being baked for ~ 48 hours at ~ 120 °C, the base pressure was typically 2x10"o Torr. The other chamber was double 1.1-metal shielded and housed a high resolution electron energy loss spectrometer (EELS) (LK Technologies LK3000). The 1.1-metal shielding is necessary to reduce the stray magnetic fields present within the chamber to < 35 mG.1 The sample was mounted at the end of a long-travel (500 mm) manipulator capable of x, y, z, and 0 motion (Thermionics 910438NW) which was connected to the spherical chamber. The sample mount was constructed in-house and is shown in Figure 2.1. The mount was connected in vacuum to the threaded end of the manipulator through a Macor block. Macor is a machinable ceramic material that provided both thermal and electrical isolation for the sample mount. Additional thermal and electrical isolation was provided by the cryobreaks (ISI 9611004). The sample was held to the molybdenum sample block by two tungsten clips which provided good thermal and 20 Top view 1 j W sample clips / \ TM bl k Macor oc Cu LN2 reservoir CUHOO) / sample Mo sample Side view mount __ m—m / 4:93:33 ee °e°\r \ Cryobreak (ISI 9611004) Heater filament holes Figure 2.1 Sample mount. mechanical contact to the mount. The M0 mount was secured to the copper liquid nitrogen (LN2) reservoir via a threaded molybdenum rod and nut. The sample temperature was monitored using an E-type thermocouple placed between one of the clips and the front face of the sample. To ensure fast response and low thermal load on the sample, the thermocouple junction was kept as small as possible (OMEGA Engineering, 0.13 mm diameter). The sample could be cooled to 85 K by drawing LN2 through the reservoir using a small diaphragm pump. The sample was heated indirectly by using a tungsten filament (Alfa Aesar, 0.25 mm diameter). The heater filament was wound (25 turns each side) around a small diameter rod and inserted into alumina tubes (Kimball Physics Al2O3-TU-C-500) prior to being inserted into the heater filament holes in the mount. In this fashion, the sample could be heated to > 800 K. All cryogen and electrical lines in vacuum were covered in braided silica sheathing and connections were made via a 2.75 in. multiport flange (MDC MMF275-5-133) attached to the manipulator. Also, the entire mount could be electrically biased via a copper wire attached to the mount. Sample Preparation. The Cu(100) sample (Monocrystal, Inc. > 99.99995 % purity) was cleaned by cycles of argon ion sputtering and annealing. The sample was heated to 680 K, at which time argon was admitted into the chamber to a pressure of 5x10’5 torr. The ion pumps were turned off and a gate valve opened to pump the chamber via a turbomolecular pump (Varian V-80) connected to a gas manifold system. The sputtering was begun with the following conditions: 1000 eV, 15 mA emission, ~ 25 uA current onto the sample mount. The sample mount geometry prevented a measure of ion current directly onto the sample. Sputtering continued for 15 minutes, at which time the ion gun was turned off and the sample was allowed to anneal at 680 K for 10 minutes prior to cooling to ~ 300 K for analysis. During the post-sputter anneal, the ion pumps were turned on when the pressure was < 1><10'6 torr. The sample was considered clean when XPS and EELS showed no contamination present. Generation of Formaldehyde and Dosing. Gas-phase H2CO (D2CO) was prepared by pyrolysis of paraformaldehyde (Aldrich, 95%) or paraformaldehyde-d2 (Aldrich, 99% atom) using a previously described method.2 Two Nupro stainless steel in-line gas filters (SS-4F-7), with the sintered filter elements removed, were packed with ~ 2 g paraformaldehyde and ~ 3 g of dried MgSO4. Both materials were constrained within the filters with glass wool that had been deactivated with trimethychlorosilane. 22 The filter assembly was attached to the gas line between the molecular leak valve and the gas manifold by a T-connector. The leak valve, gas lines, and the manifold were heated to ~ 100 °C to prevent condensation of the H2CO which was generated, along with water, by heating the filter assembly to ~ 60 °C, via heating tape. The water was trapped by the MgSO4 prior to introduction into the vacuum chamber. Residual gas analysis of the vapor introduced into the chamber indicated ~ 6 % H2O contamination. Removal of the drying agent, MgSO4, reduced the water content of the gas to ~ 3%. No effect on the chemistry of H2CO on Cu(100) by the water was observed. The majority of experiments were performed by cooling the sample to T S 85 K and backfilling the UHV chamber with H2CO for a predetermined time (background dosing). Exposures of H2CO are quoted in langmuirs (l langmuir = l L = 106 Torr . s) and are uncorrected for ion gauge sensitivity. Directed dosing was accomplished by dosing through a 1/8‘h in. OD. stainless steel tube connected to the molecular leak valve (the gasket assembly within the leak valve was threaded for 8-32 and the tube was soldered to a 8-32 bolt that had been drilled through). The sample was moved away from the end of the closer and dosing was performed by opening the leak valve, while monitoring the most abundant fragment by the QMS, until a constant pressure was achieved. The base pressure, as monitored by the ion gauge, rose by only ~ 2x10"° Torr. Once the QMS signal stabilized, the sample was moved under the closer for a predetermined time. The doser-sample distance was ~ 4 mm. Based on background dosing experiments involving H2CO, enhancement factors of ~ 300 were calculated for directed dosing using the stainless steel tube. 23 Preparation of 0/Cu(100). Oxygen-covered Cu(100) surfaces were prepared by closing dioxygen (AGA, UHP grade) through the 1/8th in. OD. stainless steel tube onto the Cu(100) at 300 K. The oxygen coverage as a function of dosing time is shown in Figure 2.2. A dose time of 300 s was used to prepare a coverage of 0.1 ML as indicated by XPS analysis. At this coverage, oxygen forms a disordered overlayer with the oxygen atoms in 4-fold hollow sites as determined by photoelectron diffraction.3 A single loss was observed in EEL spectra at 335 cm'1 (corresponding to v(Cu-O)) which matches closely with previous investigations of O/Cu(100) where a loss was observed at 340 cm'1 for a coverage of 0.11 ML.4 X-ray photoelectron spectroscopy (XPS). The surface chemical composition was determined by XPS. All XP spectra were collected using the Al Kat X-ray line (hv=1486.6 eV) operated at 300 W (15 kV, 20 mA) and an analyzer pass energy of 100 eV. Photoelectrons were collected at a take-off angle of 75° from the surface normal to maximize surface sensitivity. Spectra were referenced to the Cu (2133/2) peak from clean Cu(100) at 932.7 eV.5 Typically, photoemission data (Cu (2p), 0 (ls) and C (Is) regions) were acquired in less than 2 minutes. In order to calculate the C(15) and O(ls) XPS atomic sensitivity factors for our instrument, single point calibration experiments were performed using a saturation exposure of CO (Matheson, 99.99 %) on the clean Cu(100) surface. This system is known to form a saturated monolayer coverage of 0.57 ML (0.57 CO molecules per Cu atom) at 85 K.6 Measurement of the C (ls):Cu (2pm) XPS peak intensity ratio for a CO-saturated monolayer 'allowed us to calculate an absolute carbon atom concentration for any measured C (ls):Cu (2p3/2) 24 0.12 I 0.10 - . 0.08 - I A 0.06 - _l g XPS Analysis G O / Cu(100) 0.04 - Directed dosing of 02 0.02 - 0.00 T I"'TI""I"' 'r' 0 100 200 300 400 500 Exposure time (s) Trrfiiilrri Figure 2.2 Oxygen coverage on Cu(100), resulting from the directed dosing of 02 at 300 , as determined by XPS. 25 or O(ls):Cu (2p3/2) XPS ratio. This method was used to determine the absolute number of H2CO molecules for a given exposure. Electron Energy Loss Spectroscopy (EELS). The clean and adsorbate-covered surface was characterized by high resolution electron energy loss spectroscopy (EELS). In a typical EELS experiment, a monoenergetic beam of electrons is scattered off a surface. The majority of electrons scatter elastically, but, a small fraction interact with the surface-adsorbate system and lose quanta of energy corresponding to vibrations of the system},7 For the experiments described here, the dominant interaction was assumed to be dipolar. In this scattering mechanism, the electron scatters inelastically from the long- range dipole field produced by the adsorbate. This interaction changes the momentum minimally and, as such, the electrons that have lost energy are detected very close to the elastic peak (specular direction). Electron energy loss spectra were acquired in the specular scattering geometry (9j=03=55°) with a primary electron beam energy of 6.1 eV. The elastically scattered beam from the clean Cu(100) surface was typically 24-32 cm"1 (3-4 meV) full width at half maximum (FWHM). Under these conditions, currents onto the Cu sample were approximately 300 pA and count rates from the clean Cu(100) surface were >106 Hz. From the adsorbate-covered surface, the FWHM ranged from 32-64 cm'1 (4-8 meV) with c=0unt rates of 103-105 Hz. Temperature-Programmed Desorption (TPD). The adsorbate-covered surface was characterized by TPD. During TPD experiments, the adsorbate-covered surface was heated at a linear rate (4.2 or 5.7 K-s'l) and the desorbing molecules characterized by mass Spectrometry. The surface was positioned such that the surface normal was in 26 direct line-of-sight of the QMS. Molecules were transmitted to the mass spectrometer through a 2 mm diameter aperture in a stainless steel shroud enclosing the ionizer. The shroud ensured the majority of desorbing species originated from the surface and not the sample mount. While heating, a —70 V potential was applied to the sample to prevent electrons from the mass spectrometer ionizer inducing adsorbate chemistry. The rate of desorption for an adsorbate is equal to the rate of change of coverage (0) with respect to temperature (T) and can described by the Polyani-Wigner equation Rate of desorption = :19— : V6 exp - Ed (2.1) dT is kBT where v is the preexponential, n is the order of desorption, B is the heating rate and Ed is the energy of desorption. For zero-order desorption (n=0), Arrhenius plots of ln(rate) versus ’I“1 yield values for Ed and v. Redhead8 showed that for a first-order process Ed can be estimated from the peak desorption temperature (Tm) using equation (2.2) 13d = er[1n[v:;m ]— 3.46] (2.2) and assuming a value for v (usually 1013 5"). Equation (2.2) has been found to vary by only ~ 2 % when v falls between 108 and 10‘3 s".9 These two analyses represent only a few of the procedures cited in the literature.9'” Photochemistry. The adsorbate-covered surface was exposed to unpolarized ultraviolet (UV) radiation from a medium-pressure Hg arc lamp (Oriel 6286) operated at 350 W. The lamp was equipped with an aqueous visible/infrared filter (~ 1 M NiSO4) which transmitted photon energies in the range 5.4-3.9 eV (230-320 nm). The UVN is 27 6.19 l 1 4.96 l r Energy (eV) 4.13 r l r r 1 3.54 l I l I 3.10 0.7 - 0.6 - 0.5 - ' 0.4 - I 0.3 - Transmittance 0.2 - 0.1- I 0.0 -|—' 1 I 200 I I .,. 250 ..... 300 . Ta 350 "l 400 Wavelength (nm) Figure 2.3 Ultraviolet/Visible spectrum of ~ 1 M NiSO4 (aq) solution used as a filter for the medium-pressure Hg arc lamp. 28 spectrum of a ~ 1 M NiSO4 solution is shown in Figure 2.3. The UV light was introduced into the UHV chamber through a fused quartz window. The angle of incidence of light was 45° with respect to the surface normal. Irradiation of the chamber and sample mount was minimized through the use of an aperture affixed to the window and irradiation of the sample with a power of ~ 9 chrn‘2 resulted in a temperature rise of ~ 2 K. All irradiations were performed at T S 85 K. 2.1 References (1) Ho, W. In Investigations of Surfaces and Interfaces - Part A; Rossiter, B., Baetzold, R., Eds.; Wiley and Sons: New York, 1993; Vol. 9A. (2) Terentis, A.; Waugh, 8.; Metha, G.; Kable, S. J. Chem. Phys. 1998, 108, 3187. (3) Kittel, M.; Polcik, M.; Terborg, R.; Hoeft, J.; Baumgartel, P.; Bradshaw, A.; Toomes, R.; Kang, J.; Woodruff, D.; Pascal, M.; Lamont, C.; Rotenberg, E. Surf. Sci. 2001, 470, 311. (4) Sueyoshi, T.; Sasaki, T.; Iwasawa, Y. J. Phys. Chem. B. 1997, 101, 4648-4655. (5) Seah, M.; Smith, G. In Practical Surface Analysis; Briggs, D., Seah, M., Eds.; Wiley and Sons: New York, 1990; Vol. 1; pp 531. (6) Tracy, J. J. Chem. Phys. 1972, 56, 2748. (7) Ibach, H.; Mills, D. L. Electron Energy Loss Spectroscopy and Surface Vibrations; Academic Press: San Diego, 1982. (8) Redhead, P. A. Vacuum 1962, 12, 203. (9) de long, A.; Niemantsverdriet, J. Surf. Sci. 1990, 233, 355. (10) Koch, K.; Hunger, B.; Klepel, O.; Heuchel, M. J. Catal. 1997, 172, 187-193. 29 (11) King, D. Surf. Sci. 1975, 47, 384. 30 Chapter 3 Adsorption and Polymerization of Formaldehyde on Cu(100) Abstract The adsorption of formaldehyde (H2CO) on clean Cu( 100) at 85 K has been studied using electron energy loss spectroscopy (EELS), X-ray photoelectron spectroscopy (XPS), and temperature programmed desorption (TPD). For coverages up to 1.06 (i0.22)x1015 H2CO molecules/cmz, formaldehyde spontaneously polymerized to form a monolayer of disordered poly(oxymethylene) (POM), arranged with the chain directions parallel to the surface plane. Thermal decomposition/desorption of the polymer monolayer occurred by two routes, producing peaks in temperature programmed desorption (TPD) at approximately 200 K and 215 K. These features were produced by molecular H2CO generated via depolymerization of the polymer. The 200 K and 215 K features displayed apparent zero- and first-order desorption kinetics, corresponding to estimated activation energies for depolymerization of 75 (i 10) and 53.9 (i- 0.5) kJ/mol, respectively. The presence of two polymer desorption peaks is attributed to chain conformational differences present within the monolayer, and has not been previously observed in studies of formaldehyde adsorption on metal surfaces. Large exposures of H2CO on this surface formed multilayers of molecular formaldehyde on top of the first polymer layer. The second layer desorbed at 105 K and subsequent layers at ~100 K. 31 3.1 Introduction Compared to Langmuir-Blodgett (LB)1»2 and self-assembled monolayer (SAM)3'8 techniques, direct adsorption of monomer followed by polymerization, either thermally or through initiation with photons or electrons, offers a simple, rapid and solvent-less route to the formation of polymer thin films. However, for most of the systems which have been investigated to date, little is known about the adsorbed monolayer structure prior to polymerization, the polymer morphology or the details of the reaction mechanism. Elucidation of the polymerization initiation, propagation and termination mechanisms operative for ordered monolayers of unsaturated small molecules, and the influence of surface electronic and crystallographic structure on the polymer film order, chain length, conformation and direction(s) in relation to specific crystal directions may ultimately enable the design of high-quality crystalline polymer thin films. There have been relatively few fundamental studies of small molecule polymerization reactions on surfaces even though this type of reaction plays a large role in Fischer-Tropsch catalysis}10 While the adsorption of formaldehyde on Cu(100) has not been investigated (the focus of this chapter), there is some precedent for thermal polymerization of H2CO to poly(oxymethylene) (POM) on several metal surfaces: O/Ag(110),ll Ni(110),12 Pt(lll),13 Pd(111),14 O/Rh(111),15 O/Pd(lll),16 NiO(100)17 and Cu(110).18 Thermal polymerization of acetaldehyde on O/Ag(l 1 1) has also been observed.19 In addition to thermal polymerization, ultraviolet photons or low-energy electrons have been shown to initiate polymerization of other small molecules adsorbed at metal substrates: TCNQ,20 dinitrobenzene,21 thiophene,22v23 formaldehydez‘iv25 and styrene.26 32 3.2 Results and Discussion Temperature Programmed Desorption (TPD). The adsorption of formaldehyde on Cu(100) was studied by TPD. Figure 3.1 shows a series of m=29 (HCOI) TPD spectra for increasing exposures of H2CO (for clarity, not all exposures shown). At 0.2 L H2CO exposure, desorption peaks of approximately equal intensity were observed at 197 K and 212 K. With increasing exposure of H2CO, both peaks increased in intensity. However, the 197 K peak shifted to higher peak desorption temperatures, but the 212 K peak remained at approximately the same peak desorption temperature. The total area of these two peaks, as a function of exposure, is shown in the inset of Figure 3.1. The area increases linearly up to an exposure of approximately 0.9 L and then reaches a constant value. Separate fits of the increasing and constant regions intersected at 0.87 (10.14) L, which can be related to coverage, as discussed below. For a 1.1 L H2CO exposure, the two desorption features merged into a broad peak at approximately 205 K, and a new peak was observed at 105 K. For exposures >2.2 L, an additional peak at 100 K grew in that did not saturate with increasing exposure. Based on previous investigations,18 the 105 K and 100 K features are associated with desorption of molecular H2CO from the second and subsequent monolayers, respectively. The feature observed at 84 K is attributed to desorption from the heating filament and the large, broad background observed between 100 K and 180 K is attributed to desorption from the sample mount.27 Figure 3.2 shows TPD spectra for a 1.0 L H2CO exposure on Cu( 100) measured at masses corresponding to hydrogen (m=2, Hf), water (m=l8, H2O+), carbon monoxide 33 80 - _ (U I r- 9..) . I I < I 60 _ g ' m I r- 5 I 93 C I 40 - _ .' (ht _ 0.0 0.5 1.0 1.5 2.0 2.5 3.0 E 20 H2CO Dose (L) i E? . h a) c 9 e 0 - f s 8 r c 2 b N I l 3 macaw-'- OLIIIIJIIIILJJIIILILIII 1 00 1 50 200 250 300 Temperature (K) Figure 3.1 Mass 29 thermal desorption spectra for increasing doses of H2CO on Cu(100). All exposures were made at 85 K. The inset shows the total integrated area for the features between 190 and 230 K as a function of exposure. Exposures are: a) 0.2 L, b) 0.3 L, c) 0.5 L, d) 0.6 L, e) 0.8 L, t) 1.1 L, g) 1.8 L, h) 2.2 L, i) 2.7 L. (m=28, co”), forrnyl (m=29, Hco“), formaldehyde (m=30, H2CO+) and 13c- forrnaldehyde (m=31, H2'3CO+). Each spectrum constituted a separate 1.0 L exposure. For masses 28,29, 30 and 31, the relative abundances for each ion (20:100:48:l.3) between 190 K and 230 K were similar to those obtained from the mass spectrum of H2CO(g) (24:100:58:l.l).28 This confirms that the identity of the species desorbing between 190 K and 230 K is molecular formaldehyde. We believe molecular H2CO is generated as a result of depolymerization of a surface-bound polymer as will be discussed in detail below. For m=28, three desorption peaks were observed. The two highest temperature peaks (approximately 200 K and 215 K) can be identified as fragment ions of molecular H2CO, as described above. However, the 174 K peak was not observed at m=29, 30 or 31 and thus, is inconsistent with fragmentation of H2CO. The similarity of the desorption temperature of this peak with that of a CO-exposed Cu(100) surface (data not shown) leads us to conclude that this feature is due to desorption of molecular CO. The m=2 spectrum displays a peak at 178 K and a shoulder at ~200 K. The shoulder likely corresponds with cracking of the H2CO and is associated with comparable peaks observed at m=29, 30 and 31. However, the peak at 178 K cannot be attributed to H2CO cracking. This feature was likely due to recombinative desorption of adsorbed H atoms, which are known to adsorb on copper surfaces at 85 K2931 For the m=18 TPD spectrum, peaks are observed at 197 K and 224 K, which are associated with desorption of water coadsorbed with the formaldehyde during the gas exposure. A scan of 1-100 amu during desorption showed no other desorbing species. 35 t 1.0LHZCO/Cu(100) _ m=2 f m=18(x10) g- . 2 ' m=28 9 .- 5 . C 2 _ . m=29 M ' m=31(x50) '..r.L..r.L..r....r.r.. 100 150 200 250 300 Temperature (K) Figure 3.2 Thermal desorption spectra for masses 2 (Hf), 18 (H2O+), 28 (CO+), 29 (HCO+), 30 (H2CO+), and 31 (H213COI) for a 1.0 L exposure of H2CO on Cu(100). Each spectrum is a separate 1.0 L exposure. All exposures were made at 85 K. Spectra are offset for clarity. Electron Energy Loss Spectroscopy (EELS). Figure 3.3 shows a series of EEL spectra for increasing exposures of H2CO on the 85 K Cu(100) surface. Immediately upon adsorption of H2CO on this surface, the EELS elastic beam intensity falls by more than two orders of magnitude (from >106 Hz to ~104 Hz), indicative of a significantly disordered adlayer. Table 3.1 summarizes the EELS results together with IR data obtained from gas-phase H2CO32, solid H2CO33 and the polymer, POM.34 Excellent agreement is obtained between our EELS data for the 0.5 L H2CO-dosed Cu(100) surface and the IR data for solid poly(oxymethylene). The losses at 338 cm'1 and 2060 cm'1 can be assigned to v(Cu-CO) and v(CEO) for adsorbed CO. Irnportantly, the carbonyl loss due to v(C=O) for H2CO is not observed, suggesting no molecular H2CO is present on the surface up to about 1.2 L. However, the possibility of H2CO adsorbed with the molecular axis parallel with the surface cannot be excluded. The only mode that would be active for dipole scattering in this geometry would be the out-of—plane deformation (B2 character) which has an associated dipole moment oriented perpendicular to the surface. This mode should appear at ~1167 cm'l, which is an area of the spectrum that is dominated by polymer loss features. The presence of 1,3,5-trioxane, the cyclic trimer of H2CO, was eliminated based on EELS data taken for the adsorption of trioxane on Cu(100) at 85 K. Figure 3.4 is the comparison of EEL spectra of a 0.5 L exposure of H2CO (Figure 3.4a) with that of a 0.7 L exposure of trioxane (Figure 3.4b). Trioxane has been found to have complex adsorption 37 ,5 _ H2CO / Cu(100) 85 K 10 - 7 d 21 L ,9 x10 ) X 7,; X150 0. 8 :- .~§‘ c) 2.4 L 2 9 x250 E 5- b) 1.2 L x250 a) 0.5 L x150 IIIlllIlllJlIllIllllllllllllllllllllllLll 0 1000 2000 3000 4000 Energy loss (cm'1) Figure 3.3 EEL spectra for increasing exposures of H2CO on Cu(100) at 85 K. The sample was flashed to 300 K between exposures. 38 03m comm A; .m>+m>v 8:253. Eton 39 Sam 3% 58.; mam 2% Eu? «Sm owwm Aeamvwm £83, cowm Swm :3me EU? 88 22 5mm: 0 _ t _ E A53: Suu; NE E: mm: 83 A382 58% 82 em: 35; SS ewmtmmfl mmfl ES 3332 :29: 35:. \ 36: .Aoummvem o: _ E _ 33$: eon a: 52 8: 599g; o2: tum mmo mg 5099 Se 08 500cm mmv ocoto>o wmm A092; me 2008 852 M mm Ace—V50 v— mw Ace—EU 25$ 4 no 3.2—om 20m 255 A 3. 3.26m DUN: 280 GUN: Beeswaé .OUNI can 220% $5.ansz338 26m .8 Sun 5 2m :39? 82 .v_ mm 3 50:30 so GUN: ho 35.898 4 _m can A md .8 3mm .3 @3830 3-80 Ev mucus Econ—«.53 05 .8 mucoEchm/x _.m 2an Cu(100) 85 K 15- 8(OCO) b) 0.7 L trioxane 2: E 10 _ ‘ X150 (0 Q. 3 E? U) 1- C .92 E 0-5 " a) 0.5 L H200 x150 0.0-Jk r I uuuuuuuu I rrrrrrrr l 11111111 I rrrrr 0 1000 2000 3000 Energy loss (cm'1) Figure 3.4 EEL spectra of a) 0.5 L exposure of H2CO on Cu(100) and b) 0.7 L exposure of trioxane on Cu(100). Both spectra were acquired at 85 K. behavior on Cu(1 11) with monolayer coverage occurring at ~ 0.02 ML and multilayers forming after saturation of the monolayer.35 The EEL spectrum shown in Figure 3.4b likely corresponds to a multilayer coverage of trioxane on Cu(100) although absolute coverage was not calculated. The losses observed match well to those found for bulk crystalline trioxane as seen in Table 3.2 Also, the observation of losses corresponding Table 3.2 Assignments of the vibrational bands (in cm") observed by EELS for 0.7 L exposure of trioxane on Cu( 100) at 85 K. Also shown is IR data for solid trioxane. Assignment Symmetrya Trioxane 0.7 L Trioxane / Cu(100) Solid36 v(M-CO) 321 5(COC) A1 469 470 5(OCO) E 521 526 VS(COC) / r(CH2) A1 / E 951 / 918 946 VS(COC) E 1067 1049 Va(COC) E l 152 l 165 r(CH2) A1 1222 1215 t(CH2) E 1313 1288 w(CH2) E 1419 1392 8(HCH) E 1483 1472 v(CEO) 2059 v,(CH) E 2877 2853 _v,_,(CH) E 3031 3003 a) bulk symmetry, C 3,. to E modes in bulk trioxane (C 3,, symmetry) rules out an overlayer structure with the C3 axis parallel with the surface normal. Irnportantly, the loss observed at 645 cm'1 in Figure 3.4a, assigned to 8(OCO), most closely matches the reported value for POM at 630 cm". In the IR spectrum of solid 1,3,5-trioxane, this mode is split into two bands at 744 cm'l and 521 cm]. For the EEL spectrum shown in Figure 3.4b, the 744 cm’1 loss was undetected due to its inherently low intensity.36 41 On increasing the H2CO exposure to 2.4 L (Figure 3.3c), some recovery of the elastic beam intensity was noted (approximately an order of magnitude), indicating that the second layer is somewhat more ordered than the first (polymer) layer. The appearance of an intense loss at 141 cm'1 has been previously observed for H2CO on a number of surfaces and has been attributed to lattice (phonon) modes of crystalline H2CO(s).“‘ 1348’37’33 At 2.4 L exposure, the loss at 1483 cm'1 significantly increased in intensity and a new loss at 1719 cm“1 appeared. These losses are attributed to molecular H2CO and agree with IR data taken on crystalline H2CO films at 80 K.32 The losses at 1184 cm'1 and 1240 cm'1 also gained intensity and are assigned to the out-of-plane and in-plane bending modes of H2CO respectively. The v(C-H) region showed three distinct losses which can be assigned to vs(C-H) and vas(C-H) of H2CO and to a Fermi resonance between v2+v5 and V.;. This Fermi resonance has previously been observed in the IR spectrum of crystalline H2CO”,33 The loss due to adsorbed CO, v(CE-O), has red- shifted to 2020 cm'1 and gained intensity relative to 1.2 L, but the loss due to v(Cu-CO) was still not detected at ~340 cm". We assume the modest red-shifting of the v(CEO) mode is due to intermolecular coupling between adsorbed H2CO polymer and CO. At the largest H2CO exposure studied, 21 L, the losses due to molecularly adsorbed multilayer H2CO were observed clearly. The lattice mode has gained intensity and blue- shifted by 87 cm'1 to 228 cm"1 relative to the 2.4 L exposure. A loss at 423 cm’1 appeared and is tentatively assigned to an overtone of the 228 cm'1 lattice mode. The peak at 943 cm", for which there is no corresponding mode in molecular H2CO, is 42 assigned to the vs(OCO) mode of polymer formed during acquisition of the EELS spectrum. Previous studies of H2CO adsorption on Ag(111) have shown that polymerization of molecular formaldehyde can be initiated by low-energy electrons and that even exposure to the low fluxes of electrons encountered in EELS can initiate polymerization of multilayers of H2CO.25’39 Figure 3.5 shows a series of EELS spectra taken as a function of annealing temperature for a 6.3 L exposure (multilayer) of H2CO on Cu(100). The surface was dosed at 85 K, momentarily heated to the indicated temperature and immediately cooled to 85 K prior to data acquisition. Figure 3.5a shows the “as exposed” spectrum taken immediately after 6.3 L of H2CO on the 85 K Cu(100) surface. As expected, it displayed losses characteristic of both molecular H2CO and POM. In particular, the v(C=O) of H2CO at 1720 cm'1 and the vs(OCO) of POM at 950 cm'1 are observed. The v(CEO) loss due to adsorbed CO at 2011 cm], is observed but it is significantly red-shifted relative to CO adsorption on the bare Cu(100) surface.‘40 However, at this exposure, the loss due to v (Cu-CO) at ~340 cm'1 was not observed. Figure 3.5b shows the EELS spectrum obtained after annealing the 6.3 L exposed surface to 120 K. In contrast to the unannealed (85 K) spectrum (Figure 3.5a), the loss due to v(C=O) of multilayer H2CO was not observed and the losses due to POM were narrower and more intense (in particular the losses assigned to v(OCO) of POM), even though the full width at half-maximum (FWHM) of the elastic peaks for the two spectra were similar. We attribute this observation to increased ordering of the POM induced by 43 60 - 6.3 L (5.6 ML) H2CO/Cu(100) c) 300K "BO 40- x50 3? ’0? o. ‘0’ . b) 120K 32‘ (D E E 20‘ x50 a) 85K x50 0 rIr...rrrrrlrrrrrrrrrJJJJrrerrlrrrrrrrr. 0 1000 2000 3000 4000 Energy loss (cm'1) Figure 3.5 EEL spectra as a function of annealing temperature for a 6.3 L exposure of H2CO on Cu(100). The “as dosed” spectrum is shown in a) while b) and c) were annealed to the indicated temperature. All spectra were recorded at 85 K. The shoulder at ~ 200 cm"I on the elastic peak is the lattice mode associated with the spectrum shown in a). This peak is absent in b) and c). 44 heating to 120 K. The v(CEO) loss at 2007 cm'1 has lost intensity and the loss due to v(Cu-CO) was not observed. Annealing the sample to 300 K, as shown in Figure 3.5c, resulted in complete decomposition/desorption of the polymer as evidenced by the absence of any losses due to POM. The only losses observed are due to CO readsorbed during cooling of the sample following annealing. X-ray Photoelectron Spectroscopy (XPS). The adsorption of H2CO was also studied using XPS. Figure 3.6 shows the C (Is) and 0 (Is) regions for increasing exposures of H2CO on Cu(100) surface at 85 K. At low exposures (<1 L), a peak is observed in the C (ls) spectrum at a binding energy (BE) of 288.2 eV. With increasing exposure, this peak shifted to higher binding energy (288.8 eV for 5.4 L H2CO). Indeed, our measured C (1s) binding energy for the adsorbed polymer is similar to that measured by Pireaux et al.41 for bulk poly(oxymethylene) of 287.8 eV, suggesting that the polymer- surface interaction is weak. Similar behavior was observed for the 0 (Is) region; a peak was observed initially at 533.1 eV at 0.1 L which shifted to 533.8 eV at 5.4 L. These XPS binding energies are attributed to POM at <1 L exposure and, at higher exposures, to a mixture of POM and molecular H2CO. As mentioned above, absolute coverages of formaldehyde on Cu(100) were obtained using XPS by standardization with a known adsorbate system. Figure 3.7 displays the calculated coverage as a function of exposure for the adsorption of H2CO on Cu(100) at 85 K. From the inset of Figure 3.1, TPD data indicate saturation of the polymer desorption features occurred at 0.87 L. Using the standardized XPS intensity ratios calculated above, this exposure corresponds to an absolute coverage of 0.69 (i014) 45 538 536 534 532 530 528 526 I I I I l I I I I r I I l I I I l I I I l I T I b) 0 (1s) a) C (1s) Intensity (arb. units) 1x10‘ cps 55% M J W '1': : rI'Tn 1r1+1[L. :1 r 1 nTr'r lr—ll 1 294 292 290 288 286 284 282 280 Binding energy (eV) Figul’e 3.6 XP spectra of a) C (Is) and b) 0 (Is) for increasing exposures of H2CO on Cu( 100) at 85 K. The sample was flashed to 300 K between exposures. Exposures were 0, 0- 1 . 0.2, 0.4, 0.5, 0.7, 0.9, 1.2, 2.1, 5.4, L. Vertical lines are drawn at 288.8 eV and 533-8 eV for the C (Is) and 0 (Is) regions, respectively. 46 5 .. H2CO / Cu(100) 85 K Coverage (ML) 0 1 2 3 4 5 H200 Exposure (L) Figure 3.7 Calculated coverage as a function of exposure for the adsorption of H2CO on Cu(100) at 85 K based on XPS C(ls) peak areas. Equation of linear fit is y=0.898*x- 0.082. 47 ML (0.69 formaldehyde molecules per surface atom) or equivalent to 1.06 (3:0.22) x1015 formaldehyde molecules/cmz. Depolymerization Energetics. The adsorption and reactions of formaldehyde on the Cu(100) surface described here can be compared with previous studies of formaldehyde on clean group IB metals. On Au(110)42 and Ag(111),24 formaldehyde adsorbs non-reactively and desorbs as molecular H2CO at approximately 160 K and 110 K, respectively. In the case of Ag(111), UV or low energy electron irradiation induces polymerization, with depolymerization occurring at 210 K producing molecular H2CO. In contrast, on the 90 K Cu(110) surface, H2CO was found to adsorb reactively to form POM.18 Desorption peaks were observed at 110 K and 225 K and attributed to multilayer H2CO desorption and decomposition of the polymer, respectively. Interestingly, the depolymerization temperatures noted in our work for Cu(100) are similar to the polymer formed spontaneously on the copper (110) surface and through ' photopolymerization on the silver (111) surface (~200-230 K), implying that the polymer produced is similar in these cases. In contrast to the previous studies of H2CO adsorption on Cu(110) and Ag(111), in which a single peak due to depolymerization was observed in TPD spectra, we observe two distinct peaks. These features originate from POM. The most likely explanation for the appearance of two depolymerization features in our TPD data for Cu(100) is a difference in depolymerization kinetics and/or energetics. Expansion of the temperature region in which the depolymerization features are observed (180-240 K) (shown in Figure 3.8) reveals that the lower temperature peaks share a common leading edge and a shift to a higher peak desorption temperature with increasing H2CO exposure. For a true 48 - H2CO / Cu(100) N l Ion Intensity (m=29) Exposure Temperature (K) Figure 3.8 Enlarged view of the polymer features from the TPD spectra shown in fig. 1. A constant background was subtracted from each spectrum. Exposures ranged from 0.1 L to 1.1 L. 49 desorption process, these are the characteristics of zero-order desorption kinetics.43 However, for the experiments presented here the measured desorption rate of formaldehyde is not reflective of the desorption kinetics associated with molecular H2CO, but rather with decomposition of the polymer to produce molecular H2CO which promptly desorbs. Hence, the zero-order desorption kinetics observed are representative of the poly(oxymethylene) depolymerization mechanism. Thus, standard TPD data analyses yield activation energies that are relevant to the rate of depolymerization. The higher temperature TPD feature shows a constant peak desorption temperature, which is indicative of a first-order desorption process. We can estimate the activation energy and pre-exponential factors for depolymerization by subjecting the zero-order desorption feature to a leading edge analysis43 (plots of ln(rate) vs. T") while the activation energy for the first-order feature was calculated by the Redhead method assuming a pre-exponential of 10'3 8".44 These analyses yielded activation energies of 75 (i10) and 53.9 (10.5) kJ/mol for the zero- and first-order desorption features, respectively. These depolymerization energies fall within the broad range of 42-113 kJ/mol quoted for POM from various bulk studies.“5 The difference in activation energy is less important than the observation of two different kinetic orders for depolymerization and will be discussed in detail in Chapter 4. The most reasonable explanation for the appearance of multiple desorption peaks for polymer decomposition is the presence of different polymer species that depolymerize by a similar pathway. The possibility of a single polymer species decomposing via two competing pathways can be immediately discounted because such a situation would be 50 expected to produce a single desorption feature; the fastest and/or lowest barrier process will always dominate. Possible Polymer Species. An obvious choice for two different polymer species would be polymer adsorbed at distinct sites such as terrace sites versus. defect or step- edge sites. The appearance of two desorption peaks in the monolayer TPD data (Figure 3.1) of approximately equal intensity implies that the surface concentrations of defects and/or edge sites would be approximately equal to the concentration of terrace sites. We estimate the density of step-edges for a Cu(100) surface cut to within 0.5° (supplier’s specification) at ~l %. Furthermore, excellent surface order is suggested by the high EELS elastic peak intensity (>106 Hz) of the clean Cu(100) surface, and <20 % loss in instrumental resolution relative to the "straight-through" geometry. For these reasons, we believe defect sites and/or step-edges are not responsible for the two polymer desorption peaks. It is possible conformational differences could give rise to the two polymer TPD peaks. In the crystalline state POM forms helices belonging to either the D( 10709) or D(n') point group in the trigonal (t-POM) and orthorhombic (o-POM) crystal structures, respectively.46r47 Both point groups are isomorphous with D" and have an associated dynamic dipole oriented parallel (A2) or perpendicular (E) to the helical chain axis.34 This is shown in Figure 3.9a for orthorhombic POM. Adsorption of these helices with the primary rotation axis (z-axis in Figure 3.9a) parallel to a surface would reduce the symmetry to C1. As a consequence of the metal-surface selection rule, the original A2 modes, polarized within the surface plane, would be strictly screened and thus unobservable for dipole scattering in EELS. The E modes would still have some 51 a) DUI?) A2 I: i——- y E E Helix viewed Helix viewed along z-axis. along x-axis. b) C 2v vas(OCO), B1 VS(OCO), Al O O O z \//\\/\// 8v 0 O_ o o Figure 3.9 Comparison of symmetry for a) orthorhombic (helical) and b) planar POM. The Cartessian coordinates (x,y,z) and representations (A2, E, B1, A1) for a) correspond to the helix in free space while for b) the substrate is considered. The filled circles are methylene units (CH2). 52 component of their dipoles oriented perpendicular to the surface depending on rotational orientation. Thus, for dipole scattering, the observed features in our EELS spectra must originate from what were E modes of the polymer. Similar arguments have been made for the thermal polymerization of acetaldehyde adsorbed on O/Ag(111).19 Another chain conformation the polymer could adopt is a planar zigzag bound to the surface through the O atoms (the planar conformation has never been observed experimentally in the solid) as shown in Figure 3%. Assuming the C2,, symmetry of the planar conformation is retained upon adsorption, the vas(OCO) mode should be (strictly) screened in this conformation but observable for helical conformations. While the coexistence of the helical and planar forms cannot be ruled out on the basis of our EELS data, the observation of vas(OCO) does eliminate the possibility that only the zigzag form was present on the surface. We expect the vibrational frequencies for the planar and helical conformations of the surface-bound POM to be very similar and therefore indistinguishable by EELS. Small morphology-dependent frequency shifts have been observed between orthorhombic POM and trigonal POM.48 The largest difference observed by this study was 37 cm'1 between the o-POM (596 cm'l) and t-POM (633 cm" I) 8(OCO) mode. The vibrational frequency of the 5(OCO) mode observed in our EELS data (631 cm‘l) implies that the crystal habit present on the Cu(100) surface is trigonal. Interestingly, we have preliminary evidence that this mode red shifts to approximately 599 cm“1 upon annealing the polymer monolayer to 190 K, perhaps suggesting a transformation to the orthorhombic form (data not shown). The surface density of formaldehyde molecules in the polymer monolayer, provides additional insight into the conformation of POM on the Cu( 100) surface. It will 53 be recalled that the calculated density, based on standardized XPS measurements, was 1.06x1015 cm'z. For crystalline orthorhombic poly(oxymethylene), there are two possible planes that can be projected onto the surface to produce chain directions parallel to the surface plane (saturation of the polymer TPD features discounts polymer chain growth perpendicular to the surface plane). The bc plane has a calculated average monomer density of 7.35x10l4 cm'z, while the ac plane has a calculated average monomer density of 1.18x1015 cm'z. For trigonal poly(oxymethylene), the ac plane has an average monomer density of 1.16x1015 cm’z. Thus, it can be seen the ac plane of both o-POM and t-POM produce surface monomer densities close to that calculated by the standardized XPS data for saturation of the polymer layer. Although agreement between these surface densities is good, and suggests that the polymer monolayer covers the surface completely, it should be stressed that we have already inferred that the layer is probably highly disordered and should not be considered crystalline. The existence of a decomposition route that does not produce molecular H2CO is evident by the desorption of CO and H2 prior to the depolymerization of POM. A common origin for these two species is indicated by the coincident desorption temperatures. Leading edge analysis of these second order desorption peaks yielded an activation energy of 49 (i8) kJ/mol and a pre-exponential factor of 1x1015‘“ molecules l-cm2-s". The ratio of desorption in the H2+CO versus H2CO channels showed a maximum at ~ 0.04 at 0.66 ML (95 % of saturation of the polymer monolayer). The intensity ratio for the m=28 peak at 178 K and the m=2 peak at 177 K, corrected for ionization probability and fragmentation, is approximately 1.15, very close to the expected decomposition ratio for a species containing equal proportions of H2 and CO. 54 We have already mentioned the possibility of the existence of H2CO species bound to the surface with the molecular axis parallel to the surface plane; this adsorption geometry may give rise to the H2 and CO desorption products observed at 178 K in TPD. Therefore, the observation of H2 and CO desorption at 177 K must be due to decomposition of a third, as yet unidentified, species. This species may arise at step edge or defect site as these would be expected to more reactive than terrace sites on the Cu(100) surface. 3.3 Conclusions The adsorption of formaldehyde (H2CO) on the Cu(100) surface has been studied. At 85 K, H2CO polymerizes spontaneously to form a monolayer of poly(oxymethylene) (POM) up to a saturation coverage of 0.69 ML (1.06x10ls cm'z). This surface density suggests that the POM chain directions are parallel to the surface plane. No molecularly adsorbed H2CO was observed in EEL spectra prior to saturation. Further exposure of the polymer-covered Cu(100) surface resulted in multilayers of H2CO, with EELS loss features characteristic of crystalline H2CO. These multilayers are susceptible to polymerization by the low-energy electron beam during EELS data acquisition. Temperature-programmed desorption spectra indicate there is one decomposition route available to the adsorbed polymer species: decomposition to molecular H2CO. The route producing H2CO is observed as two desorption features at approximately 200 and 215 K which show zero- and first-order depolymerization kinetics respectively. The observation of two orders can be possibly explained by the presence of two types of polymer existing in different conformations. We believe differences due to adsorption at 55 defect and/or step edges do not account for the appearance of two desorption features for route (ii). The observation of CO and H2 desorption indicates a third species is present that is probably due to dissociation of H2CO at step edges and defect sites. Electron energy loss spectroscopy data indicate no other species besides an adsorbed polymer first layer and molecular H2CO multilayers on top of the polymer layer, are present on the 85 K Cu(100) surface. 3.4 References (1) Ejaz, M.; Yamamoto, S.; Ohno, K.; Tsujii, Y.; Fukuda, T. Macromolecules 1998, 31, 5934. (2) Shirai, E.; Urai, Y.; Itoh, K. J. Phys. Chem. B. 1998, 19, 3765. (3) Ford, J. F.; Vickers, T.; Mann, G; Schlenoff, J. Langmuir 1996, 12, 1944. (4) Sun, F.; Castner, D.; Grainger, D. Langmuir 1993, 9, 3200. (5) Kim, T.; Chan, K.; Crooks, R. J. Am. Chem. Soc. 1997, 119, 189. (6) Chan, K.; Kim, T.; Schoer, J.; Crooks, R. J. Am. Chem. Soc. 1995, 117, 5875. (7) Balasubramanian, K.; Cammarata, V. Langmuir 1996, 12, 2035. (8) Batchelder, D.; Evans, 8.; Freeman, T.; Haussling, L.; Ringsdorf, H.; Wolf, H. J. Am. Chem. Soc. 1994, 116, 1050. (9) Bent, B. E. Chem. Rev. 1996, 96, 1361. (10) Somorjai, G. Introduction to Surface Chemistry and Catalysis; Wiley and Sons: New York, 1994. (11) Stuve, E.; Madix, R.; Sexton, B. Surf. Sci. 1982, 119, 279. (12) Richter, L.; Ho, W. J. Chem. Phys. 1985, 83, 2165. (13) Henderson, M.; Mitchell, G.; White, J. Surf. Sci. 1987, 188, 206. 56 (14) Davis, J.; Barteau, M. J. Am. Chem. Soc. 1989, 111, 1782. (15) Houtman, C.; Barteau, M. Surf. Sci. 1991, 248, 57. (16) Davis, J.; Barteau, M. Surf. Sci. 1992, 268, 11. (17) Truong, C.; Wu, M.; Goodman, D. W. J. Am. Chem. Soc. 1993, 115, 3647. (18) Sexton, B.; Hughes, A.; Avery, N. Surf. Sci. 1985, 155, 366. (19) Sim, W.; Gardner, P.; King, D. J. Am. Chem. Soc. 1996, 118, 9953. (20) Wells, S.; Giergel, J .; Land, T.; Linquist, J .; Hemminger, J. Surf. Sci. 1991, 257, 129. (21) Tsai, W.; Boeri, F.; Clarson, S.; Montaudo, G. J. Roman Spectrosc. 1990, 21, 311. (22) Cheng, L.; Bocarsly, A.; Bemasek, S.; Ramanarayanan, T. Surf. Sci. 1997, 374, 357. (23) Land, T.; Hemrrringer, J. Surf. Sci. 1992, 268, 179. (24) Fleck, L.; Feehery, W.; Plummer, E.; Ying, Z.; Dai, H. J. Phys. Chem. 1991, 95, 8428. (25) Fleck, L.; Kim, J .; Dai, H.-L. Surf. Sci. 1996, 356, M17. (26) Carlo, S.; Grassian, V. Langmuir 1997, 13, 2307. (27) The large, broad desorption feature between 100 and 180 K was observed during evaluation of our experimental arrangement using a known system (CO/Cu(100)). This system generally shows no broad features in this temperature range and the appearance of features in this temperature range in our experiments was thus attributed to desorption from the sample mount. In subsequent experiments perfomed using directed dosing, these features were absent. 57 (28) NIST Chemistry WebBook, NIST Standard Reference Database Number 69; Mallard, W., Linstrom, P., Eds.; National Institutes of Standards and Technology: Gaithersburg, MD, 1998; pp (http://webbook.nist.gov). (29) Chorkendorff, 1.; Rasmussen, P. Surf. Sci. 1991, 248, 35. (30) Kammler, T.; Kuppers, J. J. Chem. Phys. 1999, 111, 8115. (31) Adsorbed hydrogen was assumed to arise from dissociation of molecular hydrogen and/or formaldehyde at the filaments of the ion gauge and/or QMS. (32) Khoshkoo, H.; Hemple, S.; Nixon, E. Spectrochim. Acta A 1974, 30, 863. (33) Weng, S.; Anderson, A.; Torrie, B. J. Raman Spectrosc. 1989, 20, 789. (34) Tadokoro, H.; Kobayashi, M.; Kawaguchi, Y.; Koybayashi, A.; Murahashi, S. J. Chem. Phys. 1963, 38, 703. (35) Hofmann, M.; Wegner, H.; Glenz, A.; Woll, C.; Grunze, M. J. Vac. Sci. Technol. A 1994, 12, 2063. (36) Kobayashi, M.; Iwamoto, R.; Tadokoro, H. J. Chem. Phys. 1966, 44, 922. (37) Anton, A.; Parmeter, J.; Weinberg, W. J. Am. Chem. Soc. 1986, 108, 1823. (38) Fleck, L.; Ying, Z.; Feehery, M.; Dai, H.-L. Surf. Sci. 1993, 296, 400. (39) Fleck, L. E. Ph.D. Thesis, University of Pennsylvania, 1994. (40) Sexton, B. Chem. Phys. Lett. 1979, 63, 451. (41) Pireaux, J.; Riga, J.; Boulanger, P.; Snauwaert, P.; Novis, Y.; Chtaib, M.; Gregoire, C.; Fally, F.; Beelen, E.; Caudano, R.; Verbist, J. J. Electron Spectrosc. Relat. Phenom. 1990, 52, 423. (42) Outka, D.; Madix, R. Surf. Sci. 1987, 179, 361. (43) de Jong, A.; Niemantsverdriet, J. Surf. Sci. 1990, 233, 355. (44) Redhead, P. A. Vacuum 1962, 12, 203. 58 (45) Muck, K. In Polymer Handbook; Brandrup, J., Irnmergut, E., Grulke, E., Abe, A., Bloch, D., Eds.; Wiley & Sons: New York, 1999; Vol. 4; pp V/97. (46) Helical symmetry is specified by the nomenclature X(2m1r/n) where X is the point group (C or D), m is the number of turns of the chain per unit cell and n is the number of monomer units per chain per unit cell). (47) Tadokoro, H. In Macromolecular Reviews; Peterlin, A., Goodman, M., Okamura, S., Zimm, 3., Mark, H., Eds.; Interscience: New York, 1967; Vol. 1; pp 119. (48) Kobayashi, M.; Sakashita, M. J. Chem. Phys. 1992, 96, 748. 59 Chapter 4 Evidence for Two Chain Length Distributions in the Thermal Polymerization of Formaldehyde on Cu(100) Abstract The polymer species formed from the spontaneous polymerization of formaldehyde (H2CO) on clean Cu(100) at 85 K were studied using electron energy loss spectroscopy (EELS) and temperature-programmed desorption (TPD). Formaldehyde forms poly(oxymethylene) (POM) with differing chain lengths; the long (or) and short (B) chain species depolymerize to give two features in TPD at approximately 207 and 219 K, respectively. The complex desorption kinetics observed for the (It-POM species were successfully modeled using equations based on the ratio of the average number of monomers unzipped from the chain per initiation event to the length of the polymer chain. Losses were observed in EEL spectra of the short chain species at ~ 290, ~ 1020 and ~ 1120 cm'1 that can be assigned to the v(Cu-O), v(C-O) and p(CH3) modes of oxygen and methoxy endgroups, respectively. Pre-adsorbed methanol increases the proportion of short chain POM species by increasing the probability of termination for the B species. Lower thermal stability for adsorbed POM was observed compared to bulk POM and is believed to be related to the stability of the surface bound oxygen endgroup. 60 4.1 Introduction In Chapter 3, investigations of the spontaneous polymerization of formaldehyde (H2CO) on clean Cu(100) at 85 K were reported.1 The formation of two types of poly(oxymethylene) (POM, -(CH2O)n-) was observed by temperature-programmed desorption (T PD) as two separate features corresponding to the depolymerization of the POM to produce H2CO(g). The lower temperature feature (~ 209 K) exhibited apparent zero-order desorption/depolymerization kinetics while the higher temperature feature (~ 220 K) exhibited first-order desorption/depolymerization kinetics.2 We ascribed the two depolymerization routes to conformational differences between the two adsorbed species. In contrast, bulk POM is known to be stable up to at least 350 K depending on the degree of polymerization and type of endgroups.3 The lower stability of the film is believed to be related to the endgroups of the polymer and has been observed previously. For example, bulk polyimide made from pyromellitic dianhydride (PMDA) and oxydianiline (ODA) is stable at temperatures up to 500 °C.4’5 However, polyimide films made from reactive adsorption of PMDA and ODA on Cu(110) decompose at ~ 280 °C.67 The thermal stability of the film was shown to be limited by the reactivity of the carboxylate endgroups bound to the surface, but the decomposition pathway and the influence of the substrate on the endgroup stability has not been thoroughly studied. In other studies on the thermal polymerization of H2CO on Cu(110), a single depolymerization feature was observed at 225 K.8 Additionally, POM formed through the photopolymerization of H2CO adsorbed on Ag(lll) depolymerized at 210 K.9 61 Unfortunately, no studies were performed to determine the kinetics of depolymerization on either surface. Regardless, the similarity of depolymerization temperatures observed for these two surfaces and the Cu(100) surface suggests the POM species formed on each surface are similar. While Chapter 3 dealt with identifying the products formed from the adsorption of H2CO on Cu(100) at 85 K, the present chapter addresses the origin of the observed depolymerization processes and the nature of the endgroups of the POM polymer. The reduced thermal stability of the POM film formed is also considered. The reaction of adsorbed methanol (CH3OH) with H2CO on Cu(100) at 85 K was used to probe the reaction mechanisms of the polymerization process and help in the identification of the POM endgroups. 4.2 Results and Discussion Isolation of POM species. The adsorption of H2CO on clean Cu(100) at 85 K resulted in the formation of two POM species. Temperature-programmed desorption was used to investigate whether the two types of POM observed to form on the Cu(100) surface could be isolated through annealing. Figure 4.1 shows m=29 (HCOI) TPD spectra over the temperature range where depolymerization of POM occurs. Figure 4.1a shows TPD data obtained from a POM-saturated Cu(100) surface (1 ML, equivalent to a fractional coverage, 0, of 0.69 H2CO molecules/Cu atom).l Two features were visible: the a-POM species with a peak desorption temperature of ~ 207 K and the B-POM species at ~ 219 K.2 Figure 4. lb shows the TPD data obtained after one monolayer of 62 HCO+ (m=29) intensity (arb. units) I I I T l 1 80 200 220 Temperature (K) Figure 4.1 Mass 29 (HCO+) TPD spectra for (a) 1 ML H2CO on Cu(100), (b) 1 ML H2CO after annealing to 209 K and cooling to 85 K prior to TPD and (c) 1 ML H2CO annealed to 209 K and dosed with H2CO to re-saturate the surface. A constant linear background was subtracted from each spectrum. 63 POM was briefly annealed (2-3 seconds) to 209 K and then cooled to 85 K prior to the TPD experiment. Annealing to this temperature, the approximate peak desorption temperature for the at species, resulted in the complete loss of the or species. Due to the sirrrilarity of peak desorption temperatures for the or and B POM, annealing also resulted in the loss of ~ 11% of the B species (based on fits discussed below). Additionally, a TPD spectrum identical to Figure 4. lb was produced if the B species-covered surface was allowed to remain at 85 K for up to 10 minutes. Figure 4.1c shows the TPD data obtained after a monolayer of POM was annealed to 209 K, cooled to 85 K, and then exposed to H2CO to re-saturate the surface with POM. The data appear qualitatively similar to that shown in Figure 4.1a. The TPD results of Figure 4.1 suggest immediately that the a species can be depopulated and repopulated by an annealing and re-dosing procedure and the B species does not convert to the or species during the TPD experiment. Irnportantly, the higher temperature B species can be isolated through annealing. Depolymerization kinetics. Equations can be derived to model the kinetics of the depolymerization process for a monodisperse polymer. For bulk POM, there are two limiting cases for depolymerization initiated at chain ends.3’10rll In the first case, once initiated, the POM chain fully depolymerizes to monomer and is first-order with respect to the mass of the polymer, W (proportional to coverage in our experiments). The rate of monomer (M) evolution, with respect to time (t), is equal to the rate of radical (R) production M =sz —— 4.1 (it dt ( ) where Z is multiple of the degree of polymerization, Dp (kap where b=1 for a monodisperse sample). The rate of radical formation is ER— : ki [11] (4.2) dt where k, is the rate constant for terminal initiation and [n] is the concentration of polymer chains. Combining equations (4.1) and (4.2), and using equation (4.3) W [n] = (4.3) MM) where Mm is the molecular weight of the monomeric unit, results in equation (4.4) EM 2 ki W (4.4) dt M The equation can be transformed into the rate of monomer evolution with respect to temperature (T) by use of the heating rate B (dT/dt), resulting in equation (4.5). M... W dT ‘BMm (4.5) The rate of monomer evolution is independent of Dp if the kinetic chain length KL (number of monomer units “unzipped” per initiation event) is equal to the degree of polymerization Dp (chain length). In the second case, following initiation, unimolecular termination competes with depolymerization and only a small number of monomers are unzipped prior to termination (KL < Dp). The rate of monomer evolution, as a function of time, can be related to the radical concentration by dM —dt_ — k d[R] (4.6) 65 where kd is the rate constant of depolymerization. The rate of radical production is dependent on the rate of initiation, k,, and the rate of termination, k, such that 95 = ki[n]- k,[R]. (4.7) (It Assuming steady state conditions for equation (4.7), using equation (4.3) and the heating rate, equation (4.6) becomes dM_ kide dT Bk,MmDp' (4.8) If the number of chains remains constant throughout the depolymerization (number of chains = W/Dp), this process is zero-order with respect to the mass of polymer. This condition would be met only during the initial stages of depolymerization as short sections of each chain are removed and both W and Dp decrease at the same rate. However, following prolonged depolymerization, the kinetic chain length (KL: kd/k,) eventually becomes comparable to Dp. At this stage, the total number of chains on the surface begins to decrease, the ratio W/Dp is no longer constant, equation (4.8) becomes first-order with respect to the mass of polymer and thus is identical to equation (4.5). Consequently, depolymerization should initially exhibit zero-order behavior (as described by equation (4.8)) but become first-order (as described by equation (4.5)) during the later stages of depolymerization. The temperature at which the apparent order changes will depend on the kinetic chain length and initial degree of polymerization and occurs when KL=Dp. Equations (4.5) and (4.8) were used to fit TPD data both as a function of heating rate and increasing coverage. Both coverage and heating rate data were fit to increase the 66 I'ITI'I'II . o O o. O .9 :::: IITIIIIIIIIlfijII'IIIIlIII I'IrIlIljl‘r I l I I I I I I I I I l I I I I l I I I I l I I I I I'I'IrI'Ilr.I .O ‘‘‘‘‘ O .... HCO+ (m=29) intensity (arb. units) I I I I I I I I I I o o u o .0 .. fI'IIII'IIII'I III I 'jjI II I I I r 1 90 200 21 0 220 230 240 Temperature (K) Figure 4.2 Mass 29 (HCO+) TPD spectra for a 1 ML coverage of H2CO on Cu( 100) exposed at 85 K as a function of heating rate (B). The heating rates were (a) 1.5 Ks'l, (b) 3.8 Ks", (c) 5.6 Ks" and (d) 8.7 Ks". The solid line (—) is the raw data, the dotted lines (...) are the fits to equations (4.5) and (4.8) and the dashed line (----) is the sum of the fits. 67 I I I I I I I l I I I I l I I I I l I I I I l I I I I L 0) 0:0.3 ML I l' I I I I I I I I I lfi I I I I I I I I I I I I I HCO+ (m=29) intensity (arb. units) ‘ . 6:0.1 ML 3) .-. ' ’1 K x1.7 o o O O '- ........ C.- uuuuuuuuuuuuuuuuuuuuuuuuuuu I I l I I I I l I I I ITrffiI I I fifil I I I I II 1 90 200 21 0 220 230 240 Temperature (K) Figure 4.3 Mass 29 (HCO+) TPD spectra for increasing coverages (0) of H2CO on Cu(100) exposed at 85 K. The coverages were (a) 0.1 ML, (b) 0.3 ML and (c) 0.6 ML. The solid line (—) is the raw data, the dotted lines (- - -) are the fits to equations (4.5) and (4.8) and the dashed line (----) is the sum of the fits. 68 reliability of the fit parameters. . Equation (4.5) alone produced a satisfactory fit to the higher temperature B-POM peak as shown in Figures 4.2 and 4.3. Since common leading edges were observed for the a—POM species as a function of coverage, equation (4.8) was used to fit the TPD data for (it-POM. However, such an approach failed to accurately reproduce the majority of the or peak shape, particularly near the peak maximum. As such, a combination of equation (4.5) and (4.8) was used to fit the or-POM species and values for k, and KJJDp determined. The rate constant kg was assumed to be of an Arrhenius form and preexponential and energetic terms extracted from the coverage and heating rate fits. Similar values cannot be determined uniquely for kd and h, only the ratio KUDp, assumed to be temperature independent in our model. The effect of varying KIJDp is illustrated in the simulated TPD spectra of Figure 4.4. When KIJDle the rate of monomer evolution follows first-order kinetics and when KIJDp<:uoE>xo;_om can 3:205 ASbESSofixon—om 23m “8 8% M: 2m :32? om? ”M mm 3 60— V50 :0 OUNQ Ea CONE ho 39238 1:2 _ e8 3mm E @3530 3.50 as momma: 232353 2: mo £=oE=m8m< NV 035‘ obtained after annealing to 194 K. The losses attributable to POM between 600 and 1500 cm’l decreased slightly in intensity while the losses at 800 and 1020 cm'1 remained approximately unchanged. Also, the S/N has increased in Figure 4.6b suggesting a modest increase in order for the annealed surface. After heating to 209 K to fully remove the or species, the spectrum shown in Figure 4.6c was obtained. Based on TPD data shown Figure 4.1, this spectrum should be entirely due to the [3 species at a fractional coverage of ~ 0.4. A new loss appeared at 280 cm", which was tentatively assigned to v(Cu-O) based on EELS studies of O/Cu(100).14 The detection of this mode was enhanced by the large SIN increase observed after annealing to 209 K, implying the [3 species was more ordered than the mixture the two species. While the losses were less intense, all modes except the strong losses at 800, 1020 and 1120 cm", were attributable to POM. The v(CH) region showed two resolved losses at approximately 2850 and 2965 cm". Annealing above 209 K resulted in the decrease of all mode intensities. Figure 4.6d shows the EEL spectrum obtained after annealing to 300 K. Complete depolymerization of both adsorbed polymer species occurred and a clean surface was recovered. The observed losses at 345 and 2087 cm'1 were due to CO re-adsorbed during sample cooling to 85 K. Electron energy loss spectroscopy, combined with the annealing studies, was also used to investigate the adsorption of formaldehyde-dz (DZCO) to determine the origin of the losses at 800, 1020, and 1120 cm". Figure 4.7 shows the adsorption of 1 ML of H2CO (DzCO) at 85 K (mixture of a and [3 species). All observed losses, except those 76 1'0 ' 1 ML H2CO / DZCO annealed to 140 K 0.8 - 92‘ 2 0.6 _ b) DZCO .9 .9 X100 P O , E e a ‘- § 0.4 - I3. 0 ‘— z ' | a) H2CO 0.2 b g ML ......... ,fi.fi.................... 0 1000 2000 3000 x50 Energy loss (cm") Figure 4.7 Electron energy loss spectra of (a) 1 ML H2CO and (b) 1 ML D2CO followed by a brief anneal to 140 K. 77 previously mentioned, can be assigned to POM (POM-d2). The losses for Figure 4.7 are summarized in Table 4.2. The vibrational frequencies obtained following DzCO adsorption (shown in Figure 4.7b) match well with IR data for solid POM- d; and the expected isotope shifts were observed”,13 Importantly, the losses at 800 and 1020 cm"1 (Figure 4.7a) were observed to shift to 760 and 945 cm'1 upon isotopic substitution, corresponding to isotopic shifts (v(POM-h2)/ v(POM-d2)) of 1.07 and 1.08, respectively. These ratios are indicative of normal modes involving primarily carbon-oxygen motion in POM, although unambiguous assignment is difficult due to the large intramolecular coupling of modes observed for solid POM. The identification of the loss for POM-h; observed at 1120 cm'1 is hampered by the coincident loss observed for POM-d2 at 1150 cm'l, assigned to vas(OCO). If the loss at 1120 cm'1 for POM-h; involved pure carbon-hydrogen motion, isotopic substitution would shift this mode, assuming an isotopic ratio of 1.4, to ~ 816 cm]. For POM-d2, this region is dominated by a strong loss at 825 cm’1 (vs(OCO)) thus obscuring the loss if present. Figure 4.8 shows EEL spectra of 1 ML of H2CO (DzCO) annealed to 209 K to fully remove the or species and isolate the B species. The results for Figure 4.8 are summarized in Table 4.2. Figure 4.8a is similar to that shown in Figure 4.6c. Upon annealing to 209 K, POM-d2 shows similar behavior to that seen for POM-hz. Namely, the losses attributable to POM-d2 decreased in intensity while the losses at 750 and 980 cm'1 (800 and 1020 cm‘1 in POM-hz) remained relatively constant. As with POM-hz, the S/N increased upon annealing and the expected isotope shifts were observed. The broad 78 1.0 - 1 ML HZCO / DZCO annealed to 209 K F 0.8 - L g; b) DZCO 9:3 0'6 _ X100 .9 U G) .u “a W E 0.4 - 0 2 0.2 . I a) H200 L x100 on J., ......... . ........ ., .. ........ . ....... 0 1000 2000 3000 Energy loss (cm'1) Figure 4.8 Electron energy loss spectra of (a) 1 ML H2CO and (b) 1 ML DZCO followed by a brief anneal to 209 K. 79 loss observed at 860 cm'1 probably arises from overlap of two losses: vs(OCO) of POM- d; and the loss at ~ 816 cm'1 (red-shifted from 1120 cm’1 for POM-hz), as was previously discussed. A loss at 271 cm"1 was tentatively assigned to v(Cu-O). The isotope shift of 1.08 observed upon deuteration suggests the loss is neither primarily derived from hydrogen (deuterium)-carbon motion nor strictly v(Cu-O) for an isolated oxygen atom. It probably arises from isotopic substitution on an atom adjacent to an oxygen atom bound to the surface. The origin of this loss as well as the other unassigned losses at 800, 1020 and 1120 cm“ will be discussed below. A possible conformation of the B-POM species can be proposed based on the EELS data in Figure 4.8. The observation of the vas(OCO) mode for both POM-h; and POM-d2, coupled with the surface selection rule, eliminates the planar zigzag conformation for the B-POM species. The most probable conformation is helical, as observed in bulk POM, and as proposed for poly(acetaldehyde) on O/Ag(111).‘5’16 Endgroup identification. The EEL spectrum shown in Figure 4.8a corresponds to the B species. If the chains are short enough, as suggested by the TPD fits, the concentration of endgroups for the B species may be sufficiently large to be visible in EELS spectra. Endgroups might be responsible for losses at 290, 800, 1020, and 1120 cm'1 for the B species which cannot be ascribed to bulk POM. Likely endgroups for POM on Cu(100) include -OH, -OCH3, and -O-Cu. The identity of the endgroups can be elucidated by comparison of the vibrational spectra of small molecules containing these functionalities adsorbed on Cu(100). For example, for monolayer methanol adsorbed on clean Cu(100), losses are observed at 610-810, 1020, 80 1130, 1450, 2850, 2945 and 3290 cm'1 that can be assigned to 8(OH), v(CO), p(CH3), 5(CH3), v(CH) and v(OH), respectively.17v18 Methoxide shows a similar spectrum, except for the lack of modes due to the hydroxyl group, and an additional loss at 295 cm’1 that can be attributed to v(Cu-O).17'19 The losses at 280, 1020 and 1120 cm", for our POM-h; (0t and B), closely match with v(Cu-O), v(CO) and p(CH3), respectively of adsorbed methanol or methoxide. Unfortunately, while all expected modes for methanol/methoxide are observed in our EELS data, they are either poorly resolved or overlap with features due to the POM backbone modes. Definitive assignment of the endgroups is also hampered by the poor S/N seen for the mixed POM overlayer. Upon annealing to 209 K to remove the or species, the losses due to the POM backbone lose intensity. This observation is consistent with the loss of longer POM chains while the shorter chains, having fewer monomer units per chain, show less intense POM losses relative to those of the endgroups. The losses at 1020 and 1120 cm", for the adsorbed POM-h; shown in Figure 4.8a, do not shift with annealing. While consistent with the v(Cu-O) of an isolated methoxide species, the mode at 280 cm’1 (B species) is, instead, assigned to a POM species bound to the surface through a terminal oxygen atom. Deuteration shifts this loss to 271 cm'1 indicating it is not an isolated oxygen atom but is likely due to the attachment of an O atom to an adjacent CH2 (CD2) group within the POM chain. For POM-d2 shown in Figure 4.8b, the loss assigned to vs(OCO) at 860 cm’1 is probably comprised of two losses: the vs(OCO) of POM- d; and the p(CD3), accounting for the apparent broadening. Based on the observed isotope shifts and the close match of 81 the losses with those of a methoxide species, we conclude the B species is short chain length POM with one end bound to the surface through an oxygen atom and the other end terminated with a methoxide group, written as -Oa-(CHzO)n-CH3, where O3 is the oxygen atom of the POM chain bound to the surface. The number of monomer units involved in the or and B species is unknown and under further investigation. The mechanism of formation of the B species will be discussed below. The loss observed at 800 cm", for POM-h; shown in Figures 4.7a and 4.8a, is more difficult to assign. Comparing the frequency with reasonable candidates, it matches best with the 5(OCO) mode of adsorbed formate observed at 758 and 780 cm‘1 on Cu(100)”,21 and Cu(110),22’23 respectively. Two characteristic losses for adsorbed formate/formic acid have been observed at ~ 1360 and ~ 1070 cm'1 that can be assigned to \I,(OCO)20v21 and 1t(CH),23 respectively. Although we observed losses at similar frequencies, the shifts upon deuteration do not match those expected for formate. The origin of 800 cm'1 loss could arise either from the bending motion, 5(OaCO), or a 150 cm' 1 red-shifted v(OaCO) mode of solid POM-h; caused by anchoring one of the oxygen atoms to the surface. The presence of isolated formate, methoxide and methanol on the Cu(100) surface can be eliminated by comparing the adsorption and desorption behavior of formic acid and methanol. Methanol exposed to clean Cu(100) has been found to adsorb molecularly and desorb at 179 K.17,18 Methoxide is formed following methanol exposure to preoxidized Cu(100) at ~ 120 K and is stable up to 400 K.17 Likewise, formic acid adsorbed on Cu(100) deprotonates to form formate. This formate species decomposes to 82 - b) CHSOH dosed first ,_ o L CH30H \ l- " 0.2 L CH30H ’3‘: .. (U E T 8 9 _ .5 I F . fi ' a) H2CO dosed first +15; - O L CH30H \ o ' 0.2 L CH30H 0 _ I ‘ “WWW . W.,, . I I L " ,l , I I ' a I f r I ' i fi 1 80 200 220 240 Temperature (K) Figure 4.9 Mass 29 (HCO+) TPD spectra after the adsorption of (a) 0.2 L methanol (CH3OH) on the Cu(100) surface pre-covered with 0.6 ML H2CO and (b) 0.6 ML H2CO on the Cu(100) surface pre-covered with 0.2 L CH30H. The “0 L CH3OH” spectrum is the same for both (a) and (b). 83 CO2 and H2 at 420 K.2031 In the present work, a clean surface was generated upon annealing to 300 K indicating no isolated methoxide or formate species were formed. Thus, isolated methoxide, methanol and formate can be eliminated as the cause of the losses observed at 290, 800, 1020 and 1120 cm'1 for the data shown in Figure 4.8a. Reaction of CH ,0H and H2CO. Since formaldehyde is known to react with methanol to form compounds of the type, CH3O-(CH2O)n-H,3 and our EELS data is generally consistent with such a species, the reaction of H2CO with CH3OH adsorbed on Cu(100) was investigated. Figure 4.9 shows TPD data obtained following the reaction of H2CO with CH3OH on Cu(100) at 85 K. Figure 4.9a displays the TPD data obtained before and after exposing a POM-covered surface to CH3OH. Upon exposure to CH3OH, little change in the TPD was observed. In contrast, when H2CO was exposed to a surface pre-covered with methanol, a dramatic change was observed in the TPD data as shown in Figure 4.9b. The proportion of the or species decreased while the B species increased. In both cases, neither a change in peak position nor width was evident. These data suggest that the POM-covered surface is non-reactive towards CH3OH but the presence of pre- adsorbed CH3OH during H2CO exposure appears to decrease the formation of long chain POM species (or). It is possible that CH3OH is controlling the preferential formation of the B species indirectly, for example by site blocking. However, since no CH3QH was observed to desorb from the surface after the exposure sequence shown in Figure 4.9b, we discount this indirect role. Most likely, CH3OH terminates the polymerization to form increased quantities of short chain POM species of the type -Oa-(CH2O)n-CH3, as previously discussed. 84 LCH3OH+HZCO/Cu(100) 1.0- 0.8— C) x150 g, L i (I) g . g 0.6- 8 b) N . ‘3 x75 E g 0.4- 0.2- . 6) JL i I x150 0.0 ..n.......,..............m...].n.... 0 1000 2000 3000 Energy loss (cm‘) Figure 4.10 Electron energy loss spectra of (a) 0.2 L CH3OH, (b) 0.6 ML H2CO dosed onto the Cu( 100) surface pre-covered with 0.2 L CH3OH and (c) the spectrum in (b) annealed to 209 K. 85 Electron energy loss spectroscopy was used to confirm that CH3OH was incorporated into the B-POM species. Figure 4.10a shows the EELS spectrum of 0.2 L CH3OH (~ 0.15 of saturation) adsorbed on Cu(100) at 85 K. The position and intensities of the observed losses match well with previous investigations for molecularly adsorbed methanol.“18 Exposing this CH3OH-covered surface to 0.4 ML H2CO results in the spectrum shown in Figure 4.10b. New losses are evident that can be assigned to POM and the resulting spectrum is similar to that shown in Figure 4.5a. Importantly, losses due to 5(OH) are absent indicating CH3OH has been deprotonated, although the detection of v(OH) is hampered by its small dipole scattering cross section.24 Also, the losses at 290, 1020 and 1120 cm", discussed previously, are evident in Figure 4.10b confirming these losses arise from the methoxide endgroup of the POM species. Annealing the overlayer shown in Figure 4.10b to 209 K results in the EEL spectrum in Figure 4.10c. Again, the data appear similar to that seen for POM formed in the absence of CH3Ol-I (Figure 4.5a). The loss at 800 cm'1 was still observed, along with those due to the methoxide endgroup and POM. Interestingly, the losses due to p(CH3) (from the endgroup) and r(CH2) (from POM) are more intense than those seen in Figure 8a. Additionally, the loss due to v(Cu-Oa) at the chain end, observed at ~ 300 cm", is also more intense. This is consistent with an increased number of chains, and thus chain ends, due to CH30H reacting with the growing chain and effectively terminating the polymerization. A reaction scheme for the formation of the B species can be proposed based on the similarities of the species formed with and without CH3OH during the adsorption and 86 polymerization of H2CO. The most likely initiation reaction is that the copper surface acts as a weak Lewis acid towards the oxygen of H2CO upon adsorption.25 Solution phase H2CO is known to polymerize in the presence of Lewis acids,3 and many oxygen containing molecules (alcohols,l7 ethers,26 acetaldehyde,27 and formic acid20»22v23) are known to interact with copper surfaces through the oxygen lone pairs. We speculate that for H2CO on Cu(100), propagation proceeds by reaction of H2CO molecules with an activated carbon atom adjacent to the surface bound oxygen. The growing chain is expected to have a lower mobility on the surface than the monomer: H2CO has been calculated to have a small binding energy of ~0.1 eV on Cu(111) and hence will be highly mobile at 85 K.23 The polymerization likely terminates by abstracting a hydrogen atom from the surface, known to adsorb on copper surfaces at 85 K.2931 This series of reactions can be written as H2CO(g) + Cu —+ Cu-O-C*H2 (4.9) CU-O-C*H2 + DCH20 —) Cll-O-CH2(OCH2)D.1-OC*H2 (4.10) Cu-O-CH2(OCH2)n-l-OC*H2 + Ha —) Cu-O—CH2(OCH2)n-1—OCH3 (4.11) where C* is the growing chain end. The initiation and termination reactions are consistent with the EELS losses observed at 290, 1020 and 1120 cm’1 that are assigned to v(Cu-O), v(CO) and p(CH3), respectively. The reactions occurring when CH30H is present during polymerization are similar to those shown above. The initiation and propagation steps shown in (4.9) and 87 (4.10) are identical but, instead of termination by a hydrogen atom, nucleophilic attack by CH3OH on the growing chain end occurs, followed by loss of a proton to the surface. This terminates the polymerization and forms the methoxy endgroup. This reaction can be written as CU-O-CH2(OCH2)n-1-OC*H2 + CH3OH3 -) CU-O-CH2(OCH2)n-OCH3 + Ha (4.12) The reactions that lead to the longer chain species (or species) are unknown at this time. A full description is limited by the poor S/N observed when both species are present and the inability to isolate the on species from the B. While the or species contains ~ 1/3 of the H2CO molecules present within the POM overlayer (at 1 ML total coverage), it is possible few chains of the longer species exist, thus lowering the likelihood that the losses from the endgroups would be observed in EELS. 4.3 Conclusions Formaldehyde polymerizes spontaneously on the 85 K Cu(100) surface forming long chain (or) and shorter chain (B) poly(oxymethylene) (POM) species. Equations describing the kinetics of depolymerization suggest the number of chains of the or species is constant for increasing coverage and determined at low coverages by a fixed number of surface initiation sites. However, the equations alone cannot explain the presence of two species based solely on differences in Dp; for equal number of monomer units, the longer chain species should depolymerize at a higher temperature. As discussed in Chapter 3, 88 conformational differences are most likely the cause of the two depolymerization features observed in TPD. Upon annealing to 209 K to fully depolymerize and remove the or species, losses are observed attributable to the endgroups of the helical B species. Losses observed at 1020 and 1120 cm'1 can be assigned to v(CO) and p(CH3), respectively, of a methoxy endgroup while a loss at 290 cm‘1 is indicative‘of the end of the POM chain bound directly to the surface through an oxygen atom. This species can be written as -Oa- (CH2O)n-CH3. A loss at 800 cm'1 is believed to arise from a mode associated with the surface-bound oxygen. Compared to bulk POM, the lower thermal stability observed for monolayer B-POM on Cu(100) most likely results from the chain being bound to the surface through the oxygen atom, resulting in a lower barrier to initiation of depolymerization. Methoxide endgroups are known to increase thermal stability in POM.3 The weakening of the POM backbone due to interaction with the surface and its influence on the thermal depolymerization is minimal, based on the similarities of the vibrational frequencies for the adsorbed and bulk POM. Pre-adsorbed methanol was found to terminate the polymerization and favor the formation of the B—POM species. The thermal stability and EEL spectra of the POM overlayer formed upon reaction with CH3OH was similar to that of B-POM formed through the direct adsorption of H2CO. However, the losses due to the endgroups were more intense than the POM layer formed by H2CO adsorption alone, indicating the B species formed were shorter than those formed through the direct adsorption process. The ability to preferentially form a specific chain length polymer with a specific 89 endgroup may open the possibility to control both thin film order and morphology and tailor thermal stability based on the interaction of the endgroup with the surface. 4.4 References (1) Bryden, T.; Garrett, S. J. Phys. Chem. B. 1999, 103, 10481. (2) The 2-3 K difference in absolute peak desorption maxima observed and reported in Chapter 4 and the previous chapter was ascribed to heating rate and thermocouple position variations. (3) Walker, J. F. Formaldehyde, 3rd. ed.; Reinhold Publishing Corp.: New York, 1964. (4) Kambe, H. In Aspects of Degradation and Stabilization of Polymers; Jellinek, H., Ed.; Elsevier: Amsterdam, 1978; pp 393. (5) Feger, C.; Franke, H. In Polyimides: Fundamentals and Applications; Ghosh, M., Mittal, K., Eds.; Marcel Dekker: New York, 1996; pp 759. (6) Haq, S.; Richardson, N. J. Phys. Chem. B. 1999, 103, 5256. (7) Plank, R.; DiNardo, J .; Vohs, J. Phys. Rev. B. 1997, 55, 10241. (8) Sexton, B.; Hughes, A.; Avery, N. Surf. Sci. 1985, 155, 366. (9) Fleck, L.; Feehery, W.; Plummer, E.; Ying, Z.; Dai, H. J. Phys. Chem. 1991, 95, 8428. (10) Reich, L.; Stivala, S. Elements of Polymer Degradation; McGraw-Hill: New York, 1971. (11) Jellinek, H. H. G. In Aspects of Degradation and Stabilization of Polymers; Jellinek, H. H. 6., Ed.; Elsevier: Amsterdam, 1978; pp 1. 90 (12) Tadokoro, H.; Kobayashi, M.; Kawaguchi, Y.; Koybayashi, A.; Murahashi, S. J. Chem. Phys. 1963, 38, 703. (13) Matsuura, H.; Murata, H. Bull. Chem. Soc. Jpn. 1982, 55, 2835. (14) Sueyoshi, T.; Sasaki, T.; Iwasawa, Y. J. Phys. Chem. B. 1997, 101, 4648-4655. (15) Tadokoro, H. In Macromolecular Reviews; Peterlin, A., Goodman, M., Okamura, S., Zimm, B., Mark, H., Eds.; Interscience: New York, 1967; Vol. 1; pp 119. (16) Sim, W.; Gardner, P.; King, D. J. Am. Chem. Soc. 1996, 118, 9953. (17) Dai, Q.; Gellman, A. J. Phys. Chem. 1993, 97, 10783. (18) Ellis, T.; Wang, H. Langmuir 1994, 10, 4083. (19) Camplin, J. P.; McCash, E. M. Surf Sci. 1996, 360, 229-241. (20) Dubois, L.; Ellis, T.; Zegarski, B.; Kevan, S. Surf. Sci. 1986, 172, 385. (21) Taylor, R; Rasmussen, P.; Ovesen, C.; Stoltze, P.; Chorkendorff, I. Surf. Sci. 1992, 261, 191. (22) Bowker, M.; Haq, S.; Holroyd, R.; Parlett, P.; Poulston, S.; Richardson, N. J. Chem. Soc., Faraday Trans. 1996, 92, 4683-4686. (23) Hayden, B.; Prince, K.; Woodruff, D.; Bradshaw, A. Surf. Sci. 1983, 133, 589. (24) Ibach, H.; Mills, D. L. Electron Energy Loss Spectroscopy and Surface Vibrations; Academic Press: San Diego, 1982. (25) Fleck, L.; Howe, P.; Kim, J.; Dai, H. J. Phys. Chem. 1996, 100, 8011. (26) Meyers, J .; Street, 8.; Thompson, S.; Gellman, A. Langmuir 1996, 12, 1511. (27) Lamont, C.; Stenzel, W.; Conrad, H.; Bradshaw, A. J. Electron Spectrosc. Relat. Phenom. 1993, 64/65, 287-296. (28) J. Greeley and M. Mavrikakis, to be published. 91 ‘\. a (29) Adsorbed hydrogen was assumed to arise from dissociation of molecular hydrogen and/or formaldehyde at the filaments of the ion gauge and/or QMS. (30) Chorkendorff, I.; Rasmussen, P. Surf Sci. 1991, 248, 35. (31) Kammler, T.; Kuppers, J. J. Chem. Phys. 1999, 111, 8115. 92 Chapter 5 Photochemistry of Formaldehyde Adsorbed on CO-saturated Cu( 100) Abstract The photochemistry of formaldehyde (H2CO) adsorbed on CO-saturated Cu(100) at 85 K was studied using electron energy loss spectroscopy (EELS) and temperature- programmed desorption (TPD). Formaldehyde was weakly adsorbed on CO/Cu(100) and desorbed at 104 K, corresponding to a desorption energy of 18.2 (i0.8) kJ/mol. Irradiation of the H2CO/CO/Cu(100) surface caused the molecularly adsorbed H2CO to polymerize, forming poly(oxymethylene) (POM). Irradiation also caused the formation of ethylene glycol (CH2OH)2. Losses observed at 870 and 3365 cm", after UV irradiation, were assigned to v(CC) and v(OH) modes, respectively, of (CH2OH)2 indicating ethylene glycol was formed promptly upon irradiation. The presence of (CH2OH)2 was confirmed by studying the adsorption of (CH2OH)2 on clean and oxygen- covered Cu(100). The formation of ethylene glycol was likely governed by geometric constraints present within the formaldehyde overlayer. 93 5.1 Introduction Formaldehyde polymerizes spontaneously to form poly(oxymethylene) (POM), - (H2CO),.-, upon adsorption on a variety of clean and oxygenated surfaces: O/Ag(110),l Ni(110),2 Pt(lll),3 Pd(111),4 O/Rh(111),S O/Pd(111),6 NiO(100),7 Cu(110)8 and Cu(100).9 Although the polymer has been positively identified through a variety of techniques, in most cases little is known about the precise initiation, propagation or termination mechanisms that lead to the polymer. The morphology, crystallinity, conformation and chain length of the polymer produced are similarly poorly characterized. Recently, formaldehyde molecularly adsorbed on Ag(111) has been polymerized to POM using photons and electrons, offering the potential for precise temporal and spatial control of creation and deposition of polymer monolayers.10~ll Such control also increases the prospects for understanding surface polymerization reactions at a more fundamental level. We recently reported the facile thermal polymerization of formaldehyde on Cu(100) at 85 K.9 The film appeared to be composed of two different types of polymer chain with different conformation and/or chain length as inferred from vibrational spectroscopy and thermal desorption/depolymerization measurements.12 We turn our attention here to the photopolymerization of formaldehyde and the nature of the polymeric film produced. In this case we use carbon monoxide as a spacer layer to prevent direct interaction of the formaldehyde with the Cu(100) surface and thereby inhibit spontaneous polymerization. Co-adsorbed carbon monoxide (CO) is known to influence the chemistry observed on surfaces. For example, cosadsorbed CO has been found to increase the 94 stability of ethylidyne on Ru(001),13 decrease the stability of the methyl hydrogen atoms of toluene on Ru(001),l4 perturb the decomposition pathway of methylamine on Ru(001)15 and promote the decomposition of saturated hydrocarbons on Ni(755).16 Carbon monoxide is weakly chemisorbed on Cu(100), desorbing molecularly at ~ 180 K.9 This desorption temperature is some 20-40 K lower than the decomposition temperature of the POM polymer but about 80 K higher than the desorption temperature of the formaldehyde monomer on CO/Cu( 100). Such a situation offers the possibility for forming patterned polymer films through photopolymerization, followed by desorption of monomer (from the unirradiated areas) and the CO spacer layer. In this way, it may be possible to selectively deposit intact polymer onto specific regions of the Cu surface in a controlled fashion. 5.2 Results and Discussion Photochemistry of H2CO/C0/Cu(100). The adsorption and photochemistry of formaldehyde (H2CO) on CO-saturated Cu(100) (CO/Cu(100)) was investigated using temperature-programmed desorption. At 85 K, H2CO adsorbs molecularly and desorbs as a single peak at 104 K as shown in Figure 5.1a (0 min.). This feature does not saturate with increasing coverage and spectra for increasing coverages have coincident leading edges (data not shown), suggestive of zero-order desorption. An Arrhenius plot of ln(rate) vs. T1 indicated a desorption energy of 18.2 (:08) lemol, confirming H2CO is physisorbed on CO/Cu(100). This desorption energy compares favorably to that seen 95 =30 x1 ' m=30 =34 x400 Irradiation time 30 min. Irradiation . 15 min. Ion mtensrty (arb untls) . , . ' l >§ - time . 5 min. A “I 5 min. 15 min. 0 min. 30 min. ¥ . i I I I I I I I j l I I 80 100 120 140 200 300 400 Temperature (K) 'rjIIlIIIIITi Figure 5.1 Temperature-programmed desorption spectra for (a) m=30 (H2CO+) and (b) m=30 (H2CO+) and m=34 (D2COD+) for a 1.1 ML coverage of H2CO (D2CO) on C0- saturated Cu(100) as a function of UV irradiation time. D2CO was used to reduce coincident mass fragment interferences. 96 for H2CO physisorbed on Ag(111) of 25 lemol.10 Importantly, no desorption features were observed between 200-240 K where the polymer, formed through the thermal polymerization on the clean Cu(100) surface, was found to depolymerize.9 This immediately suggests the CO-saturated surface is inert towards thermal polymerization at 85 K and the CO inhibits the polymerization possibly through a site—blocking mechanism. No other species were observed to desorb from the unirradiated surface except for the CO monolayer, which desorbs at ~ 180 K for a saturation coverage. Figure 5.1 also shows TPD spectra for a 1.1 ML H2CO on CO/Cu(100) as a function of irradiation time. Irradiating the overlayer for 5 minutes caused a decrease in the molecularly adsorbed H2CO intensity as observed in Figure 5.1a for m=30 (H2CO+). Continued irradiation caused this feature to decrease further. The peak desorption temperature was constant and the peak broadened slightly for increasing irradiation times. New features were observed for m=30 after 5 minutes irradiation at ~220 and ~ 235 K as shown in Figure 5.1b. These two features increase in intensity, as a function of irradiation time, and merge into a large peak centered at ~ 230 K with a small shoulder at ~ 240 K. Measurement of other fragments associated with H2CO (HCO+ and H213CO+) resulted in identical TPD spectra for the features at 104 K and between 220-250 K and whose intensity ratios reflected that of the mass spectral fragmentation pattern for gas- phase H2CO.l7 This confirms the features observed at 104 and between 220-250 K are due to molecular H2CO. Importantly, comparison of pre- and post-irradiation TPD spectra for CO (m=28), indicated no change in the desorption temperature or the coverage of the saturated CO layer. 97 Molecular formaldehyde has been observed to desorb ~ 205 and ~ 220 K due to the depolymerization of poly(oxymethylene) (POM) formed from the thermal reaction on clean Cu(100).9 The desorption temperatures observed following irradiation in the current work were similar to those observed for the depolymerization of POM formed through the photopolymerization of H2CO on Ag(111).10 This suggest the features observed between 220-250 K for m=30, as shown in Figure 5.1b, are due to the depolymerization of POM formed upon photopolymerization of H2CO adsorbed on CO/Cu(100). The presence of POM will be confirmed using EELS data as discussed below. The slight shift to higher depolymerization temperatures, observed in Figure 5.1b, could be related to changes in chain length or endgroup stability and is currently under investigation. A desorption feature was observed for m=31 at 350 K that did not show any corresponding intensity at m=29 or 30 (data not shown). This feature is inconsistent with molecular H2CO. To remove interference from coincident mass fragments and help in the identification of this peak, the adsorption and photochemistry of D2CO on CO/Cu(100) was investigated. Identical TPD data to H2CO were obtained for D2CO for the features at 104 K and between 220-250 K after irradiation. The peak observed at 350 K for m=31 (H2CO) was detected at m=34 for D2CO, as shown in Figure 5.1b, indicating the fragment contains at least 3 protons. A number of small molecules can be eliminated as the source of the 350 K desorption feature. The lack of intensity for m=29 excludes adsorbed alkoxides as the source of the 350 K feature. These species are known to desorb at temperatures 2 350 K on Cu(100),18 however, at these temperatures, alkoxides undergo B-hydride elimination 98 to produce the corresponding aldehydes, all of which show m=29 as the most abundant mass fragment in their mass spectrum.17 Also, the lack of intensity at m=60, coupled with lack of intensity at m=29, eliminates the simplest dialdehyde, ethanedial (CDO)2, as the source of the feature at 350 K. The most likely species that gives rise to the feature at 350 K is ethylene glycol. This is consistent with previous experiments on the adsorption of ethylene glycol on O/Cu(110)19 and O/Ag(110)20 where desorption features were observed at 390 and 365 K, respectively. In contrast to the present work, ethylene glycol was observed to undergo B-hydride elimination to produce ethanedial on both O/Cu(110) and O/Ag(110). It is possible, without surface oxygen present, molecular desorption is favored over ethanedial generation for the ethylene glycol produced in the irradiation of H2CO adsorbed on CO/Cu(100). The presence of ethylene glycol will be confirmed by the EELS data discussed below. The amount of ethylene glycol formed saturates after 15 minutes of irradiation, as shown in Figure 5.1b, and was calculated using both TPD and XPS data. Initially, 1.1 ML of H2CO was adsorbed on CO/Cu(100). After 15 minutes of irradiation, ~ 80 % of the original 1.1 ML H2CO desorbed as monomer (0.45 ML) or formed polymer (0.43 ML), leaving ~ 20 % unaccounted for H2CO. The lack of knowledge regarding the electron impact ionization cross-section of ethylene glycol prevented quantification using TPD. However, XPS analysis of a sample that had been irradiated for 15 minutes and then annealed to 270 K to remove monomer, polymer and CO, revealed a coverage of 0.17 ML of carbon, consistent with the remaining 20 %. 99 1.1 ML H2CO + 0.57 ML CO/Cu(100) ' d)15 min. - “1 A 450K ' x100 r E . 'E 3 — 9' g . .4? m .. 5 x20 x300 .E ' E - b)15 min. E WW 0 z ”20 x100 a)0min. 85K x4 x20 x150 l""l""l""l"" 0 1 000 2000 3000 4000 Energy loss (cm'1) Figure 5.2 Electron energy loss spectra for a 1.1 ML coverage of H2CO on CO-saturated Cu(100) for (a) 0 minutes irradiation at 85 K, (b) 15 minutes irradiation at 85 K, (c) 15 minutes irradiation followed by an anneal to 270 K and (d) 15 minutes irradiation followed by an anneal to 450 K. Each spectrum constituted a separate 1.1 ML coverage of H2CO. The data were acquired at 85 K. The adsorption and photochemistry of H2CO on CO/Cu(100) was investigated using EELS. Figures 5.2a and 5.2b show EEL spectra for 1.1 ML H2CO adsorbed on CO/Cu(100) after 0 and 15 minute irradiation. Figure 5.2a shows 1.1 ML H2CO on CO/Cu(100) prior to irradiation. The losses observed at 350 and 2080 cm'1 are due to v(Cu-CO) and v(CO), respectively, of the adsorbed CO and are observed at these approximate energies for all the spectra shown in Figure 5.2. The other losses were assigned to molecular H2CO and match well with IR data for crystalline formaldehyde.21 The frequencies are only slightly shifted from the solid-phase data confirming H2CO was weakly adsorbed on the CO/Cu(100) surface. The observation of lattice modes at 200 and 230 cm’1 suggests the adsorbed H2CO was well ordered. Also, the observation of all the modes of molecular H2CO indicates the molecule was adsorbed with the C-0 bond axis tilted from the surface normal. This geometry would render the 13. (v4, v5) and B2 (v6) modes within the C2,. point group, to which H2CO belongs, visible according to the surface selection rule.22 The data from Figure 5.2a is summarized in Table 5.1. After 15 minutes of UV irradiation, the spectrum shown. in Figure 5.2b was obtained. New modes appeared at 600, 950, 1120, 1480 and 2930 cm'1 that can be assigned to the polymer, POM, and match well with both the IR data for solid POM23 and POM formed through the photopolymerization of H2CO adsorbed on Ag(111).lo The data from Figure 5.2b is summarized in Table 5.1. The observation of the mode at 600 cm'1 is characteristic of the 5(OCO) mode of POM and eliminates trioxane (cyclic trimer of H2CO) as a photoproduct. The corresponding mode for trioxane is split into two bands which appear at 744 and 521 cm".24 101 @0388 8: n E DUNE 82883390an E0: 20m 3 102 23 59> cam 83 3 3+3 28% Emma 33 £83, 22% Emma 28. EB; 8% 3823 58...; 8% e352 £0? $8 88 6mm; 8: 3,: E Gnu: 8.2 82 cm: :2 3582 50:2 22 35? owe e332 .6 o: _ 3&2 _ saw 8. _ o. _ 2 52 609...; oma 9a N8 509; ohm coo coo one 609m 9% cm 60-22 CMN DUOE 033w— ONN OUOE 033m— .E .56 2 .E .55 o 22.8 52.8 82 55598”: 82 EQOQOUN: .2 A _ 2 33:22 22 Comm anaemia. .2 _ Sw< :0 20m 5m 8% 3mm 5:5 mac—w 20m can 00$.— 2.8 .5“ Sat Mn 2a :32? 82 .cocfithm >3 8558 2 can o Sam 53 V30 93823-00 co CONE mo omab>8 1:2 2 a c8 notomno A780 .5 855 352223 05 Co mucoficwEm/x _.m 035. Interestingly, the losses due to molecular H2CO, in particular the loss at ~1720 cm'1 for v(C=O), were not observed even though, based on TPD measurements, it makes up ~ 40 % of the overlayer after 15 minutes of irradiation. The most probable explanation for the absence of these modes is a change of orientation, rendering the modes inactive according to the surface selection rule, coupled with the reduced signal- to—noise (SIN) observed in Figure 5.2b. The absence of the strong phonon losses seen in Figure 5.2a indicates the H2CO present after irradiation was no longer well ordered. The spectrum in Figure 5.2b appears qualitatively different than that seen for the thermal polymerization of H2CO on clean Cu(100).9 The losses observed in the thermal polymerization at 1220 (r(CH2)), 1390 (w(CH2)), and 1470 (8(HCH)) cm'1 are much more intense than those seen for the present work. As seen in Figure 5.2b, losses at 1220 and 1390 cm'1 were not detected. Additionally, the loss observed at 1030 cm'1 for the thermally polymerized POM was also absent. This loss has previously been identified as a mode (v(CO)) due to an endgroup of the POM chain.12 The differences in intensity could be due to orientation effects and/or chain length dependence (the absence of the endgroup mode in the present work implies longer chains). There are two losses in Figure 5.2b that cannot be assigned to POM. The small shoulder observed at 870 cm'1 is in the correct range for a mode due to v(C-C) while the broad loss centered at 3365 cm'1 is probably due to v(OH). Adsorbed water could give rise to the loss at 3365 cm", however, this loss was not detected prior to irradiation as shown in Figure 5.2a. More likely, these two loses were due to the feature observed to desorb at 350 K in the TPD data that was tentatively assigned to ethylene glycol. 103 "I; ' Annealing the overlayer shown in Figure 5.2b to 270 K resulted in the spectrum shown in Figure 5.2c. This temperature was high enough to desorb the monomeric H2CO (104 K), CO (180 K) and the POM (220-250 K) with the remaining losses due to the feature observed to desorb at 350 K. The broad loss at 890 cm“1 matches well to that seen in Figure 5.2b suggesting the species desorbing at 350 K was formed upon irradiation and not during the temperature ramp for the TPD experiment. Another loss was observed at 1070 cm'1 which most likely corresponded to a mode involving carbon-oxygen motion. The v(OH) mode at 3365 cm'1 was lost upon annealing and no v(CH) modes were detected. The spectrum in Figure 5.2c matches well with that observed for deprotonated ethylene glycol on O/Ag(110) where losses were observed at 890 and 1090 cm’I corresponding to v(CC) and v(CO), respectively.20 Therefore, based on TPD and EELS data, the species observed to desorb at 350 K was identified as ethylene glycol. Annealing the sample to 450 K resulted in the spectrum shown in Figure 5.2d. All losses due to molecular H2CO, POM and ethylene glycol were absent; the species formed during irradiation had desorbed. The observed losses were due to CO re- adsorbed during sample cooling. Ethylene glycol on Cu( 100) and 0/Cu( 100). The adsorption of ethylene glycol (CH2OH)2 on clean and oxygen-covered Cu(100) was studied using TPD. Figure 5.3 shows m=31 (CH2OH+) spectra for three exposures of (CH2OH)2 on clean Cu(100). A single desorption feature was observed at 237 K for a 0.1 L exposure. For increasing CXposures, this peak grew in intensity and the peak desorption temperature remained unchanged. Above 0.1 L, a new feature was observed at 220 K that neither saturated nor shifted with increasing exposures. Measurement of other mass fragments of ethylene _ (CHZOH)2/Cu(100) 70‘ _ ,é‘ _ 0.5L(CHZOH)2+ :55 g ? O/Cu(100) :3 .93 ' - .E E _ .5 3 1- .? (III) a) " N E E g ' .............. :1 300 350 400 450 F"): Temperature (K) N n E E: L ON I . 0 Increasing ' exposure I ' ' . ' I T . ' ' I ' 1 ' r I ' ' 100 200 300 400 Temperature (K) Figure 5.3 Temperature-programmed desorption spectra for m=31 (CH2OH+) following exposures of 0.1 L, 0.2 L and 0.5 L of ethylene glycol (CH2OH)2 on clean Cu(100). The inset is the TPD spectrum obtained for a 0.5 L exposure on a Cu(100) pre-covered with 0.1 ML oxygen showing a new feature desorbing at 347 K. 105 glycol confirmed (CHZOH)2 adsorbed and desorbed molecularly. The features at 220 and 237 K are assigned to multi- and monolayer states, respectively. These temperatures match well with those seen for (CH20H); adsorption on clean Ag(l 10) where multi- and monolayer states were found to desorb at 205 and 225 K, respectively.20 Molecularly adsorbed (CHZOH)2 was found to desorb at ~ 220 K from the Cu(110) surface,19 consistent with the current results. Assuming first-order kinetics for the state at 237 K on Cu(100) and a preexponential factor of 1013 s", a desorption energy of 59 kJ/mol was estimated. This value compares favorably to the 60 kJ/mol desorption energy found for the monolayer state on Ag(110).20 The inset of Figure 5.3 shows the m=31 TPD spectrum of a 0.5 L exposure of (CH20H)2 on a Cu(100) surface pre-covered with 0.1 ML oxygen. The desorption of the multi- and monolayer states were unaffected by the oxygen. A new feature was observed at 347 K which corresponds to desorption of (CH20H)2 that had reacted with the adsorbed oxygen forming a dialkoxide species. It is known that alcohols, including ethylene glycol, adsorb molecularly on oxygen-covered copper and silver surfaces at temperatures < 120 K.18'20 Heating the sample above ~ 200 K causes the alcohols to deprotonate by reaction with the adsorbed oxygen, desorbing as water and forming a surface-bound alkoxide species. These species are stable to temperatures 2 350 K where they desorb molecularly in competition with production of the corresponding aldehyde. The result shown in the inset of Figure 5.3, showing desorption at 347 K, is consistent with previous experiments on O/Cu(110)19 and O/Ag(110)20 where this dialkoxide state was found to desorb at 395 and 365 K, respectively. Importantly, the temperature observed for the desorbing dialkoxide species matches closely with that for 106 la! 1-4 " b) (CHZOH)2 + 0.1 ML O/Cu(100) 1.2 - E .E . 3 ' 1.0 - E, 0.5 I. g ' Annealed 270 K 515 0.8 — X100 .5 'O .. .‘é’ a 0.6 - g a) (CHZOH)2 / Cu(100) 2 I 0.4 h . 0.1 L 0.2 - , 85 K . x100 0.0 ] L 'IIIIIIIIf'IIfIIIIIUrII 0 1 000 2000 3000 4000 Energy loss (cm'1) Figure 5.4 Electron energy loss spectra for (a) 0.1 L exposure of ethylene glycol (CH20H)2 on clean Cu(100) and (b) a 0.5 L exposure of (CH20H)2 on a Cu(100) pre- covered with 0.1 ML of oxygen followed by annealing to 270 K. 107 the feature observed to desorb at 350 K after 15 minutes of UV irradiation of 1.1 ML on CO/Cu(100), supporting the identification of this species as ethylene glycol. The adsorption of (CH20H); on clean and O/Cu(100) was also studied using EELS. Figure 5.4a shows the EEL spectrum of a 0.1 L exposure (CH20H)2 on clean Cu(100) at 85 K. This exposure only populated the monolayer state as seen in Figure 5.3. The losses observed match well with liquid-phase IR data25 for (CH20H); and that observed for (CH20H); adsorbed on clean Ag(1 10).20 The loss observed at 2060 cm’1 was due to adsorbed CO. The results of Figure 5.4 are summarized in Table 5.2. Noticeably absent from the monolayer spectrum was the loss due to vas(CO). This indicates the (CH20H)2 was most likely bound to the surface via both oxygen atoms in a bidentate configuration with the O-H bonds parallel with the surface and the C-0 bonds perpendicular to the surface. This geometry would have the effect of rendering inactive the vas(CO) mode as well as increasing the intensity of the mode due to r(OH). Similar arguments have been made for the monolayer spectrum of (CH20H)2 adsorbed on clean Ag(l 10).20 Annealing a 0.5 L dose of (CH20H)2 on a Cu(100) surface pre-covered with 0.1 ML of oxygen to 270 K resulted in the spectrum shown in Figure 5.4b. At this temperature the multi- and monolayer states have desorbed and a fraction of the total adsorbed (CH20H)2 has been deprotonated through reaction with the oxygen. X-ray photoelectron spectroscopic analysis indicates a coverage of 0.1 ML of (CHZOH)2 remaining. The losses due to “C(OH) and v(OH) at 700 and 3320 cm", respectively were absent in Figure 5.4b confirming deprotonation via reaction with the adsorbed oxygen. 108 "T 33588 5: n E M 05m @2355 £53 :5on G 36on oExoxEQ 3 x53 30on Am 8mm 88 gm 85: 28$ 23% 283 288m 38 58...; 288 28% 28% 28% EN So? 28 88 omom 88 28 amp; 2: _ 82 _ ow: 83; am: E38 3: 85 m8: 255% 88 88 mm: 35? NE 35: sou: 2882 82: 82: 882 28. 52 So? 222 :82: mm: 68.; 28$ 28wa 28mm 28% eww 353 com 82 zoom 28% 283 .88 $8 65: o: o: 88 £0: 8% ME. 6588 88m 65: new 0.5.582 .8355 0.5.552 .8355 8:535 .835 32%: 83555556 +2558 A8355555£ +3523 :25qu 55:3 35:3 22:57? AGO—EEO co NAIONEUV ES 2: Cw<\O co NAIONEUV 5% 83 mqmm 5:3 mac? 2:083an 2:5: .8 83 M: 8m :32? 82 .8230 58833-00 co OUNI 50 $3025 42 _._ a be M Ohm 8 321353 new 253335 >3 883:: m. wage—.8 33035 86on 2t 88 ATEo a: 3:3 3:285; 05 no mason—383‘ N.m 2an 109 The losses remaining at 1090 and 2880 cm'1 can be assigned to vs(CO) and v(CH) modes, respectively, of the adsorbed alkoxide. The loss at 880 cm'1 was comprised of two modes, v(CC) and p(CH2). This assignment matches well with that for the alkoxide on O/Ag(110) where the corresponding modes were observed at 890, 1090 and 2860 cm", respectively.20 In contrast to the work on O/Ag(110), the losses due to w(CH2) (1340 cm") and 8(HCH) (1450 cm") were not positively identified on the O/Cu(100) surface although there appears to be a weak, broad loss in that range. This is probably due to the lower surface coverage used in the present work, which resulted in a lower S/N. Again, the absence of the vas(CO) mode indicates the alkoxide species was bound to the surface through the oxygen atoms in an upright configuration. The results of Figure 5.4b can be compared with the EEL spectrum obtained for the species observed to desorb at 350 K after 15 minutes of UV irradiation of 1.1 ML H2CO on CO/Cu(100). Figure 5.5a shows the spectrum of the species formed upon irradiation and is identical to that shown in Figure 5.2c. Figure 5.5b shows the EEL spectrum of the alkoxide species formed through the adsorption of (CH20H)2 on O/Cu(100). Excellent agreement was observed between the spectra. The losses due to v(CC)+p(CH2) and vs(CO) were observed at 880 and 1090 cm", respectively for the adsorbed alkoxide and at 890 and 1070 cm‘1 for the species produced during the irradiation as shown in Figure 5.5a. These spectra confirm the species observed in Figure 5.5a as ethylene glycol. The same experiments shown in Figure 5.5 were repeated with the deuterated species: DzCO on CO/Cu(100) and (CD20H)2 on O/Cu(100). The spectra obtained are shown in Figure 5 .6. For the irradiation and subsequent annealing of D2CO on 110 _ b) 0.5 L (CH20H)2 + 0.1 ML O/Cu(100) O ' 8 E S ’ Anneal 270 K g " o 8 7.; 8 a 2 ' x50 '93 . .E 8 .. g a) 1.1 ML HZCO + 0.57 ML CO/Cu(100) m n E E O o z — i " 8 - 0° 15 min. irr. Anneal 270 K - o 8 . L ‘ . “‘ x600 x20 l""lfi"fil""lrfi 0 1000 2000 3000 Energy loss (cm") Figure 5.5 Electron energy loss spectra for (a) 1.1 ML coverage of H2CO on C0- saturated Cu(100) irradiated for 15 minutes followed by an anneal to 270 K and (b) a 0.5 L exposure of (CH20H)2 on a Cu(100) pre-covered with 0.1 ML of oxygen followed by an anneal to 270 K. 111 b) 0.5 L (CDZOH)2 + 0.1 ML O/Cu(100) ' 8 $2 — O 8 8 g .2 2 a S - 8 Annea|270K Q l\ 3 — 8 g 5 ‘3 ' x200 £3 .E 8 .. g _ a) 1.1 ML 02cc + 0.57 ML CO/Cu(100) E 2 o 0 Z - 2" °’ 8 . 8 :2 °° 15 min. irr. _ E Anneal27OK O N . L x600 x20 I""I"'rI'*'FI" 0 1000 2000 3000 Energy loss (cm") Figure 5.6 Electron energy loss spectra for (a) 1.1 ML coverage of DzCO on C0- saturated Cu(100) irradiated for 15 rninut L exposure of (CD20H)2 on a Cu(100) p an anneal to 270 K. es followed by an anneal to 270 K and (b) a 0.5 re-covered with 0.1 ML of oxygen followed by 112 CO/Cu(100), as shown in Figure 5.6a, losses were observed at 800, 970, 1070, and 1180 cm”1 that can be assigned, through comparison with Figure 5.6b, to v(CC), w(CH2), vs,a,(CO) and 5(HCH), respectively. Excellent agreement was again observed between the spectra. The results of Figure 5.5 and 5.6 are summarized in Table 5.2 along with liquid IR data of (CH20H)2 and (CH20H)2 adsorbed on O/Ag(110). From the above TPD and EELS data, the results of the adsorption and photochemistry of H2CO adsorbed on CO/Cu(100) can be now summarized. Formaldehyde physisorbed on the CO-saturated surface at 85 K desorbs molecularly at 104 K. Upon UV irradiation, POM was formed that was observed to depolymerize, and desorb as H2CO, between 220 and 250 K. A second photoproduct was observed to desorb at 350 K that was identified as ethylene glycol. The formation of both POM and (CHZOH)2 will be discussed below. Formation of POM. The polymerization of H2CO could proceed via three photoexcitation mechanisms. A thermal mechanism is discounted based on the observation of an insignificant 2 K temperature rise upon irradiation. Gaseous H2CO has a S1 (— 80 (1t* 6— n) band origin at 3.49 eV26 which is unlikely to be perturbed upon physisorption on CO/Cu(100). Ultraviolet irradiation, in the wavelength range used here, can dissociate H2CO via two channels: H2CO —> H2 + CO (molecular) and H2CO —> H + HCO (radical). The molecular channel has its threshold at 3.52 eV.26 However, polymerization is not expected to be initiated by the molecular photoproducts and no H2, which is not known to adsorb on Cu(100) at 85 K, was observed to evolve upon irradiation. The radical channel, whose threshold is at 3.73 eV,26 may initiate the polymerization through reaction of the radical species with molecular H2CO. However, 113 for experiments performed on amorphous H2CO films deposited on C5], no polymerization occurred upon irradiation with 4.0 eV photons.27 For both channels, dissociation is expected to compete with electronic relaxation which is known to occur on metal surfaces on a time scale of ~ 10'13 — 10'[4 5.23 The effect of the CO spacer layer will be to decrease the relaxation rate. The most probable excitation mechanism leading to POM is that proposed for H2CO photopolymerization on Ag(111).10’29v3O In this mechanism, absorption of the UV irradiation by the substrate produces hot electrons, which then populate the 1t* state of H2CO, forming H2CO'. This radical anion then dissociates into products that can initiate the polymerization. In the gas-phase, H2CO' dissociates into H’ + HCO31 while on Ag(l 1 1), H2CO” was observed to dissociate into CH2 and 0.29 The dissociation products on CO/Cu(100) are not yet known. In contrast to the work on Ag(111), where subvacuum electrons were produced upon irradiation, the photon energies used in the current work will produce both subvacuum and free photoelectrons. The work function of a CO—saturated Cu(100) surface was found to be 4.39 eV32 and the majority of intensity from the lamp used in this work (~ 80 %) occurs between 5.5 and 4.9 eV, indicating most of the electrons produced will be free, low energy photoelectrons. Formation of Ethylene Glycol. The formation of ethylene glycol can be explained based on the well known photochemistry of carbonyl compounds.”34 These molecules are known to be excellent hydrogen atom abstractors in their singlet 1(n1t"‘) and triplet 3(IlTl:"‘) electronically excited states, with the majority of chemistry occurring on the triplet surface. Formaldehyde, in its triplet state, can be thought of a biradical which abstracts a Proton from another molecule to form a hydroxymethylene radical 'CHZOH. 114 F I I. For the UV irradiation of H2CO on CO/Cu(100), excited formaldehyde, CHzOa abstracts a proton from a neighboring H2CO forming 'CHZOH and 'HCO. The formyl radical can then initiate the polymerization of H2CO to form POM. The hydroxymethylene radical can react in two ways: 'CHzOH could couple to another 'CHZOH, forming ethylene glycol (CHZOH)2 directly or it could react with another CHzO to form 'OCH2CHZOH. This radical species could then abstract a hydrogen from another H2CO to form (CH20H)2 and another formyl radical. The saturation of the formation of ethylene glycol after 15 minutes of irradiation may be due to geometric constraints. Ab initio results on hydrogen abstraction from methane by formaldehyde indicates only approach of the hydrogen within the molecular plane of H2CO results in reaction and an upper limit on the barrier was estimated to be ~ 77 k.I/mol.35»36 This barrier was strongly dependent on Cuzco-Corr, separation with the distance equal to 2.53 A at the transition state.35 Once ethylene glycol is formed and POM polymerization initiated by the formyl radicals, the remaining electronically excited H2CO may be in an unfavorable geometric position within the overlayer to abstract further hydrogens from either H2CO, POM or ethylene glycol. Unfortunately, the structure of the physisorbed H2CO on CO/Cu(100) is unknown at this time. However, the structure of crystalline formaldehyde does have pairs of H2CO molecules that are oriented correctly for H-atom abstraction to occur.37 The formation of ethylene glycol through the reaction of electronically excited formaldehyde has not been observed in studies of solid phase H2CO”,38 nor in the photopolymerization of H2CO on Ag(111).10 The presence of the CO spacer layer could influence the photochemistry by lengthening the lifetime of the excited formaldehyde 115 In"; 'Jo . a), (H2CO‘) by moving it away from the metallic surface or through the formation of a unique formaldehyde overlayer geometry that is favorable to the hydrogen abstraction reaction. 5.3 Conclusions The presence of a saturation coverage of CO on Cu(100) inhibited the thermal polymerization of H2CO which formed a weakly adsorbed overlayer on CO/Cu(100). Upon 30 minutes of UV irradiation, H2CO formed POM (4O %) and ethylene glycol (20 %). The photopolymerized POM depolymerized at higher temperatures than those observed for POM formed from the thermal polymerization on clean Cu(100), suggesting a longer chain length or more thermally stable endgroup. Additionally, no losses due to endgroups were observed in EEL spectra taken after irradiation. The initiation of polymerization likely occurs via the radical anion H2CO' formed from electron capture of photoelectrons generated from irradiation of the CO/Cu(100) surface. The use of CO to block the thermal polymerization of H2CO on Cu(100) and subsequent Photopolymerization shows the potential for controlling polymer thin film formation for 1‘ eactive monomer-substrate combinations by use of a co-adsorbate. The polymer film can be formed in a “controlled” fashion using photons and can then be deposited on the SUbstrate after annealing to remove the co-adsorbate. The formation of ethylene glycol from the irradiation of H2CO adsorbed on CO/CuUOO) has not been observed previously in studies of solid phase H2CO nor in the photopolymerization of H2CO on surfaces. The formation likely occurs via electronically e"(Cited H2CO which abstracts a proton from a neighboring H2CO molecule forming a 116 hydroxymethylene radical °CHzOH which then reacts with other °CHzOH radicals or H2CO ultimately forming (CHZOH)2. The saturation of formation of ethylene glycol likely results from geometric constraints present within the overlayer. 5.4 References (l) Stuve, E.; Madix, R.; Sexton, B. Surf. Sci. 1982, 119, 279. (2) Richter, L.; Ho, W. J. Chem. Phys. 1985, 83, 2165. ( 3) Henderson, M.; Mitchell, G.; White, J. Surf Sci. 1987, I88, 206. (4) Davis, J.; Barteau, M. J. Am. Chem. Soc. 1989, 111, 1782. (S) Houtman, C.; Barteau, M. Surf. Sci. 1991, 248, 57. (6) Davis, J.; Barteau, M. Surf. Sci. 1992, 268, 11. (7) Truong, C.; Wu, M.; Goodman, D. W. J. Am. Chem. Soc. 1993, 115, 3647. (8) Sexton, 3.; Hughes, A.; Avery, N. Surf. Sci. 1985, 155, 366. ( 9) Bryden, T.; Garrett, S. J. Phys. Chem. B. 1999, 103, 10481. ( l O) Fleck, L.; Feehery, W.; Plummer, E.; Ying, Z.; Dai, H. J. Phys. Chem. 1991, 95, 8428. ( 1 1 ) Fleck, L.; Kim, J.; Dai, H.-L. Surf. Sci. 1996, 356, L417. ( l 2) Bryden, T.; Garrett, S. Langmuir , in press. ( l 3 ) Sasaki, T.; Kawada, F.; Aruga, T.; Iwasawa, Y. Surf. Sci. 1992, 278, 291. ( 1 4) Rauscher, H.; Menzel, D. Surf. Sci. 1995, 342, 155. ( l 5) Sasaki, T.; Aruga, T.; Kuroda, H.; Iwasawa, Y. Surf. Sci. 1992, 276, 69. (1 6) Orita, H.; Kondoh, H.; Nozoye, H. Chem. Phys. Lett. 1994, 228, 385. 117 "L 'JJ 'J.) (17) NIST Chemistry WebBook, NIST Standard Reference Database Number 69; Mallard, W., Linstrom, P., Eds.; National Institutes of Standards and Technology: Gaithersburg, MD, 1998; pp (http://webbook.nist.gov). (18) Dai, Q.; Gellman, A. J. Phys. Chem. 1993, 97, 10783. ( 19) Bowker, M.; Madix, R. Surf. Sci. 1982, 116, 549. (20) Capote, A.; Madix, R. J. Am. Chem. Soc. 1989, 111, 3570. (21) Khoshkoo, H.; Hemple, S.; Nixon, E. Spectrochim. Acta A 1974, 30, 863. (22) Ibach, H.; Mills, D. L. Electron Energy Loss Spectroscopy and Surface Vibrations; Academic Press: San Diego, 1982. (23) Tadokoro, H.; Kobayashi, M.; Kawaguchi, Y.; Koybayashi, A.; Murahashi, S. J. Chem. Phys. 1963, 38, 703. (24) Kobayashi, M.; Iwamoto, R.; Tadokoro, H. J. Chem. Phys. 1966, 44, 922. (25) Sawodny, W.; Niedenzu, K.; Dawson, J. Spectrochim. Acta A 1967, 23A, 799. (26) Moore, C.; Weisshar, J. Ann. Rev. Phys. Chem. 1983, 34, 525. (27) Mansueto, E.; Ju, C.; Wight, C. J. Phys. Chem. 1989, 93, 2143-2147. (28) Zhou, X.; Zhu, X.; White, J. Surf. Sci. Rep. 1991, I3, 73. (29) Fleck, L.; Howe, P.; Kim, J .; Dai, H. J. Phys. Chem. 1996, 100, 8011. (30) Fleck, L. E. Ph.D. Thesis, University of Pennsylvania, 1994. (3 1 ) Azria, R. Ph.D. Thesis, University of Orsay, 1972. (32) Dubois, L.; Zegarski, B. Chem. Phys. Lett. 1985, 120, 537. (33) Turro, N. In Molecular Photochemistry; W. A. Benjamin, Inc.: New York, 1965; PD 1 37, 118 (34) Klessinger, M.; Michl, J. In Excited States and Photochemistry of Organic Molecules; VCH Publishers: New York, 1995; pp 395. (35) Severance, D.; Pandey, B.; Morrison, H. J. Am. Chem. Soc. 1987, 109, 3231. (36) Sumathi, K.; Chandra, A. J. Photochem. Photobiol. A:Chem. 1988, 43, 313. (37) Weng, S.; Torrie, B.; Powell, B. Mol. Phys. 1989, 68, 25. (38) Gol’danskii, V. Ann. Rev. Phys. Chem. 1976, 27, 85. 119 p33}: tfllfl ”1".“ 5L; at let‘s ml: iii:- a:— 75’ 100 011 it Chapter 6 Conclusions and Future Work 6.1 Thermal Reactions and Control of Stability The work described in Chapter 3 has shown that at 85 K, H2CO spontaneously polymerizes to form a monolayer of poly(oxymethylene) (POM) up to a saturation coverage of 0.69 ML (1.06x1015 cm'z). This surface density suggests that the POM chain directions are parallel to the surface plane and is consistent with the known crystal structures of POM.1 However, the POM overlayer is probably not highly ordered. Temperature-programmed desorption spectra indicate there is only one decomposition route available to the adsorbed polymer species: depolymerization to molecular H2CO. The route producing H2CO is observed as two desorption features at approximately 200 and 215 K which show apparent zero- and first-order depolymerization kinetics, respectively. This behavior has neither been observed for H2CO adsorption Cu(110)2 nor for the POM produced in the photopolymerization of H2CO on Ag(111).3'6 We believe differences due to adsorption at defect and/or step edges do not account for the appearance of two depolymerization features. The Observation of CO and H2 desorption indicates H2CO is probably dissociating at step edgfits and defect sites upon adsorption. Electron energy loss spectroscopy data indicate "0 Other species besides an adsorbed polymer first layer and molecular H2CO multilayers on top of the polymer layer, are present on the 85 K Cu(100) surface. The observation of two features for the depolymerization of POM on Cu(100) can be aSCribed to the formation of long chain (on) and short chain ([3) poly(oxymethylene) (POM) species as described in Chapter 4. Equations describing the kinetics of 120 f-.. F b 6 1 I) s 1.. n. I a.“ E . W. .E )1 “is 5P depolymerization suggest the number of chains of the CL species is constant for increasing coverage and determined at low coverages by a fixed number of surface initiation sites. Future experiments should include increasing the number of defect sites, which may increase the proportion of a—POM species produced upon adsorption of H2CO if initiated at these sites. Chapter 4 has demonstrated that equations describing the bulk depolymerization process can be used to model the TPD spectra which should be applicable to future surface polymerization studies where multiple depolymerization processes are observed Upon removal of the (it species by annealing, losses are observed attributable to the endgroups of the helical B species. Losses can be assigned to v(Cu-O), v(CO) and p(CH3) of the end of the POM chain bound directly to the surface through an oxygen atom and to a methoxy endgroup. This species can be written as -Oa-(CHzO)n-CH3. Compared to bulk POM, the lower thermal stability observed for monolayer B-POM on Cu( 100) most likely results from the chain being bound to the surface through the oxygen atom, resulting in a lower barrier to initiation of depolymerization. We believe, due to Similarities in depolymerization temperature and product (molecular H2CO), the a—POM SPeCies likely has similar endgoups to B-POM. Pre-adsorbed methanol was found to terminate the polymerization and favor the fol‘Ination of the B-POM species. The thermal stability and EEL spectra of the POM ovePlayer formed upon reaction with CH3OH were similar to that of POM formed throllgh the direct adsorption of H2CO. However, the losses due to the endgroups were mot e intense than in the POM layer formed by H2CO adsorption alone, indicating the B s - . . p ecles formed were shorter than those formed through the direct adsorption process. 121 The experiments with the co-adsorbed methanol and DzCO also lend support to the directly bound oxygen controlling the thermal stability of the POM species. For bulk POM, methoxy termination is known to be more stable than the hydroxy endgroup.7 This indicates the oxygen bound end is likely responsible for the ~ 100 K lower stability observed for POM on Cu( 100) as compared to bulk POM. The ability to preferentially form a specific chain length polymer with a specific endgroup may open the possibility to control both thin film order and morphology and tailor thermal stability based on the interaction of the endgroup with the surface. Transforming the oxygen bound end into a methoxy terminus should increase the thermal stability dramatically. Poly(oxymethylene) dimethyl ethers are known to be stable to temperatures > 470 K, some 100 K higher than the dihydroxy terminated paraformaldehyde.7 A potentially simple way to form a methoxy endgroup from the surface bound oxygen is through the reaction with methyl radicals (CH3°). The methyl radicals can be generated photochemically from co-adsorbed CH3Br8 or dosed from the gas phase as CH3', produced from the pyrolysis of azomethane.9 Methyl radicals are known to be highly mobile, even at 85 K, on Cu(111)10 and thus could react with the OXYgen bound to the surface forming a methoxy endgroup. However, the CH3' may also deStroy the POM backbone chain through hydrogen abstraction reactions. As in the methanol experiments, the order of adsorption may influence the products greatly. Acetoxy—terminated bulk POM is also known to be more thermally stable than the hydr(My-terminated form.7 On the Cu(100) surface, acetoxy endgroups may be formed by Dre-adsorbing acetic acid prior to adsorption of H2CO. In a similar fashion to the forlhation of methoxy endgroups with pre-adsorbed methanol, described in Chapter 4, 122 on C iii} 1 "P 131:... L acetoxy endgroups would thus be formed. In this case, as acetic acid is more acidic than methanol, the initiation and termination mechanisms for polymerization may be different and both endgroups may now be acetoxy. Further experiments with co-adsorbates may allow for a full elucidation of the propagation and termination events during the surface polymerization process. 6.2 Photochemical Reactions UV and X-ray Photons. The presence of a saturation coverage of CO on Cu(100) inhibited the thermal polymerization of H2CO which formed a weakly adsorbed overlayer on CO/Cu(100). Upon irradiation for 30 minutes with UV photons, H2CO formed POM (40 %) and ethylene glycol (20 %). The photopolymerized POM depolymerized at higher temperatures than those observed for POM formed from the thermal polymerization on clean Cu(100). Additionally, no losses due to endgroups were observed in EEL spectra taken after irradiation, consistent with longer chains. The initiation of polymerization likely occurs via the radical anion H2CO" formed from electron capture of photoelectrons generated from irradiation of the CO/Cu(100) surface. X-ray irradiation also caused polymerization of H2CO adsorbed on CO/Cu(lOO), Izlll‘ther supporting a substrate mediated process. Figure 6.1 shows TPD spectra displaying the effect of irradiating 1.1 ML of H2CO on CO/Cu(100) with X-ray photons (hv=1486.6 eV, re=os nm). Figure 6.1a shows the m=29 (HCO+) TPD prior to irradiation and is identical to Figure 5.1a-b (0 min.). A single feature at 104 K is observed corresponding to molecularly adsorbed H2CO. Figure 6.1b shows irradiation of the oVerlayer for 84 s 123 1.1' ML HzCO / CO-sat / Cu(100) f I c)1303 " , b)84s HCO+ (m=29) ion intensity (arb. units) . , . . , 1 00 200 300 400 Temperature (K) Figul‘e 6.1 Mass=29 (HCO+) TPD spectra showing the effect of X- -ray irradiation on 1.1 I--I‘12CO on CO/Cu(100). Irradiation times were a) 0, b) 84 and c) 130 seconds. The 5933)’ wavelength was the Al KOt (hv=1486. 6 eV) line and operated at 300 W (20 mA,15 124 SEC the is? 10 causes the feature at 104 K to decrease while a new feature is observed at 228 K that can be identified as H2CO arising from the depolymerization of POM formed upon irradiation. Further irradiation causes a further decrease in the feature at 104 K, an increase in the H2CO due to POM depolymerization and a shift to higher peak depolymerization maximum (233 K), as shown in Figure 6.1c. The large number of secondary electrons produced from the substrate upon X-ray irradiationll likely initiates the polymerization via the radical anion channel, although more work is necessary to elucidate the mechanisms of polymerization. The ability to form a polymer thin film using X-ray photons is advantageous in the field of lithography where feature sizes of < l ()0 nm are needed.”14 New Co-adsorbates. While the use of CO is convenient, its desorption temperature (~ 180 K) prevents a determination of the “true” thermal stability of the POM formed in the irradiation of H2CO adsorbed on CO/Cu(100) because, after the CO has desorbed, the POM then can interact with the substrate, lowering the depolymerization temperature as was discussed in Chapter 4. A more suitable co- adsorbate would be one that is stable on Cu(100) above the bulk POM depolymerization temperature of ~ 370 K. An excellent candidate for this is CH3O(ad) formed from the reaction of methanol on a oxygen covered Cu(100) surface. This species forms a c(2x2) overlayer (GCH30=O.5) similar to the CO/Cu(100) overlayer and is stable up to 400 K at Which temperature B-hydride elimination occurs and H2CO desorbs”,16 In this fashion, the stability of the POM may be deconvoluted from its interaction with the substrate. The C1‘130 covered Cu(100) surface may also be prepared from methyl nitrite (CH3ONO), e1ilhinating the need for the oxygen precoverage.17 125 . 1.1 ML H2CO/Cu(100) - b) 15 min. UV irradiation GB a) 0 min. UV irradiation H200+ (m=30) Ion mtensnty (arb units) . >1 T l ' l ' 1 I. i i I I r I I I I I 250 300 . . . . , 150 200 Temperature (K) Figure 6.2 Mass=30 (H2CO+) TPD spectra for the UV irradiation of 1.1 ML H2CO adsorbed on a surface that had been pre-covered with a saturation of poly(oxymethylene) (0.69 ML) for a) 0 minutes and b) 15 nrinutes UV irradiation. 126 Preliminary work has been done using another co—adsorbate system. Formaldehyde molecularly adsorbed on a POM saturated Cu(100) surface does polymerize upon UV irradiation. Figure 6.2 displays TPD spectra (a) prior to and (b) after 15 minutes of UV irradiation of a 1.1 ML coverage of molecularly adsorbed H2CO on a POM saturated Cu(100) surface. There are three features observed prior to irradiation which can ascribed to desorption of physisorbed H2CO (105 K) and H2CO arising from depolymerization of the or and B-POM species at ~ 210 and ~ 215 K, respectively. Fifteen minutes of UV irradiation causes both the physisorbed H2CO and oc-POM species to decrease and a new feature to appear at 234 K. The feature at 234 K is consistent with a new POM species and is similar to that observed for the photopolymerization of H2CO on CO/Cu(100), shown in Figure 5. lb, where a feature is observed at 235 K. The explanation for the decrease in (it-POM species, while little Change in the B-POM species was observed, is still under investigation. However, the results shown in Figure 6.2b do suggest other co-adsorbate systems can be used to produce a more thermally stable POM thin film. 6.3 References ( 1) Tadokoro, H. In Macromolecular Reviews; Peterlin, A., Goodman, M., Okamura, 3., Zimm, B., Mark, H., Eds.; Interscience: New York, 1967; Vol. 1; pp 119. (2) Sexton, B.; Hughes, A.; Avery, N. Surf. Sci. 1985, 155, 366. (3) Fleck, L.; Feehery, W.; Plummer, 13.; Ying, 2.; Dai, H. J. Phys. Chem. 1991, 95, 8428. (4) Fleck, L.; Ying, z.; Dai, H.-L. J. Vac. Sci. Technol. A 1993, 11, 1942. 127 (5) Fleck, L.; Kim, J .; Dai, H.-L. Surf. Sci. 1996, 356, L417. (6) Fleck, L.; Howe, P.; Kim, J .; Dai, H. J. Phys. Chem. 1996, 100, 8011. (7) Walker, J. F. Formaldehyde, 3rd. ed.; Reinhold Publishing Corp.: New York, 1964. (8) Lamont, C.; Conrad, H.; Bradshaw, A. Surf. Sci. 1993, 280, 79. (9) Bent, B. E. Chem. Rev. 1996, 96, 1361. (10) Chan, Y.; Chuang, P.; Chuang, T. J. Vac. Sci. Technol. A 1998, 16, 1023. (l 1) Briggs, D.; Riviere, J. In Practical Surface Analysis; Briggs, D., Seah, M., Eds.; Wiley and Sons: New York, 1990; Vol. l-Auger and X-ray Photoelectron Spectroscopy; pp85 (12) Lawes, R. Appl. Surf. Sci. 2000, 154-155, 519. (13) Wallraff, G.; Hinsberg, W. Chem. Rev. 1999, 99, 1801. (14) Cerrina, F. J. Phys. D: Appl. Phys. 2000, 33, R103. (1 5) Dai, Q.; Gellman, A. J. Phys. Chem. 1993, 97, 10783. (16) Andersson, S.; Persson, M. Phys. Rev. B. 1981, 24, 3659. (17) Ihm, H.; Scheer, K.; Celio, H.; White, J. Langmuir 2001, I 7, 786. 128 1"“ APPENDICES 129 "'"A D? Ax“ Appendix A Mass Spectra Prior to use in experiments, the molecules studied in this thesis were first analyzed for purity. This was accomplished by opening the molecular leak valve to introduce some of the gas-phase molecules into the UHV chamber. The mass spectrum was acquired after a steady state pressure (as measured by the ion gauge) was achieved within the chamber. Typical operating condition of the mass spectrometer were 1-50 amu mass range, 10 amu/s scan rate, and 1.7 kV electron multiplier voltage. 130 3.0 h 2.5 _ b) H200 - -9 2-0 '- Pchamber ~ 1X10 TOIT ~i 1.5 — 1.0 1 a) Background Partial Pressure (Torr) x10'10 (0 C 2.5 _ .10 _ Pcmmr ~ 2x10 Torr 2.0 - 1.5 — 1.0 - “L h. .1, . .fi . 0 1O 20 30 40 50 m/z (amu) Figure A.1 Mass spectra of a) background vacuum prior to introduction of H2CO and b) after a pressure of ~ 1x10'9 Torr of H2CO was achieved within the chamber. 131 1.4 1.2 - 1.0 - b) D200 -9 o 0.8 — P chambep 1x10 Torr E 0.6 - X 5 0.4 - t W s 02 J. [fl/11.4% 931.4 a. %1-2 - a) Background 0- 1,0 _ Pchamber ~ 2x10'1o Torr 0.8 - 0.6 - 0.4 0.2 I I I I jfi I I I I I I f I I I I I I I I IiI r 0 10 20 30 40 50 m/z (amu) Figure A.2 Mass spectra of a) background vacuum prior to introduction of DzCO and b) after a pressure of ~ 1x10'9 Torr of D2CO was achieved within the chamber. 132 100 80 - _ b) Trioxane -9 60 - me3r ~ 5x10 Torr 8 N O IL IIIIIIII'IIII'IIIIIIIII'IIII[III Partial Pressure (Torr) x10'9 1o 8 a) Background _ ~10 . Pchamw ~ 2x10 Torr 61-1- 4. 2 . JUM'W WWW IIIIIIIII'ITII'IIIIIIIII'IIIll—rIIIITTTIIIllI'lIII O 10 20 30 40 50 60 7O 80 90 100 m/z(amu) Figure A.3 Mass spectra of a) background vacuum prior to introduction of trioxane (C3H603) and b) after a pressure of ~ 5x 10'9 Torr of trioxane was achreved wrthrn the chamber. 133 1.0 0.8 0.6 .-‘ .° 9 o N 1: Partial pressure (Torr) x10'10 9 CD 0.6 0.4 0.2 b) Ethylene Glycol b ' -9 . Pcmber ~ 1x10 Torr I I I r I r I r I l I I I I ‘r I I r I I I fi I I " a) Background P -10 POWnber ~ 2x10 Torr r I I I I I I I I I I I I I I l I I I 20 30 4O m/z (amu) Figure A.4 Mass spectra of a) background vacuum prior to introduction of ethylene glycol ((CH20H)2) and b) after a pressure of ~ 1x10'9 Torr of ethylene glycol was achieved within the chamber. 134 Appendix B Calculation of Electron Impact Ionization Cross-Sections To calculate the purity of a compound accurately, based on the mass spectra presented in Appendix A, and quantitate surface species from TPD spectra, the electron impact (EI) ionization cross-section (GE!) needs to be known. While the experimental 0'51 is known for a few of the compounds of interest (H2O, H2 and CO),1v2 the 0’51 for H2CO is unknown to this author. However, binary-encounter—Bethe (BEB) theory has been shown to predict (within 10-15%) the CE] for small molecules using molecular orbital constants easily obtained from standard ab initio electronic structure programs such as GAMESSJ'3 The BBB theory combines the Mott description for hard electron-electron collisions with that of the dipole interaction theory that accounts for electron interaction at high incident electron energies. The theory provide a simple, analytic expression for the 0'51 per molecular orbital (MO) and the total om is then a sum over all MOs. Three orbital constants are needed: the binding energy B, the orbital kinetic energy U and the electron occupation number N. The expression, as a function of incident electron energy T,is: GE]: 5 1n(t)1——13-+1—-1———l“(t) (3.1) t+u+1 2 t 1 11+] where T t=—- B.2 B ( ) U u=— B.3 B ( ) 135 _ 47ra(2,NR2 S B2 (B4) with a0=0.5292 A and R=13.61 ev. The orbital constants were calculated at the minimized geometry at the Hartree- Fock level with a 6-31G* basis set using the GAMESS4 code. The orbital constants for H20 and H2CO are summarized in Table BI and the 051 as a function of T is shown in Figure B.1. The GE. and orbital constants for H2O have been previously calculated and are compared to those calculated for this work as a measure of the error. The variation arises due to differences in basis set used (6—311G vs. 6-31G*) and the use of the experimental ionization energy for the HOMO of water in the previous work.2 All quantitation is done using calculated 0'51 to eliminate large discrepancies that may arise between experimental and calculated values. 136 Table B.1 Molecular orbital constants calculated at the Hartree-Fock level of theory with a 6-31G* basis set. These constants were used to evaluate the electron impact ionization cross-section. H2O“ Molecular Orbital Binding Energy (eV) Kinetic Energy (eV) 2a] 36.46 70.14 1b2 19.12 47.78 3a] 15.47 58.73 1b. 13.52 61.89 H20b Molecular Orbital Binding Energy (eV) Kinetic Energy (eV) 2a. 36.88 70.71 1b2 19.83 48.36 3a] 15.57 59.52 1b] 12.61 61.91 H2CO Molecular Orbital Binding Energy (eV) Kinetic Energy (eV) Bay 38.56 73.68 4a1 23.59 44.02 1b2 18.95 36.77 5a. 17.74 61.17 1b; 14.69 47.46 2b2 11.83 54.58 a) this work b) Reference 2 137 4 . III-III... * I. 'II I III... 1 I I. . . "i 3- I 1 I g . ‘gt‘ ‘0 6 2’2L‘ I I s . e m A b . ' A .l I - ‘c‘ I HZCO 1- A 0 H20 (this work) ' A H20 (reference 2) : fi..1.......rs,r... 50 100 150 200 Electron Energy, T (eV) Figure B] Electron impact ionization cross-section (051) as a function of incident electron energy for H2CO and H20 and a comparison to previous work done for water showing the small error introduced by using a smaller basis set. 138 B.1 References (l) Beran, J .; Kevan, L. J. Phys. Chem. 1969, 73, 3866. (2) Hwang, W.; Kim, Y.; Rudd, M. J. Chem. Phys. 1996, 104, 2956. (3) Kim, Y.; Rudd, M. Phys. Rev. A. 1994, 50, 3954. (4) Schmidt, M. W.; Baldridge, K. K.; Boatz, J. A.; Elbert, S. T.; Gordon, M. S.; Jensen, J. H.; Koseki, S.; Matsunaga, N.; Nguyen, K. A.; Su, 8. J.; Windus, T. L.; Dupuis, M.; Montgomery, J. A. J. Comput. Chem. 1993, 14, 1347. 139 Appendix C Electron Energy Loss Spectrometer and Operating Voltages Shown in Figure C1 is a schematic of the ELS3000 electron energy loss spectrometer manufactured by LK Technologies, Inc. The operating voltages for the individual segments are shown in Table C.1. These voltages were used as initial parameters prior to tuning the spectrometer in the “straight through” geometry. The resolution obtained using these voltages was 3.2 meV (26 cm") full width half maximum (FWHM) and provided ~ 70 pA of current detected at the channeltron cone. 140 Table C.l Operating voltages for the ELS3000 electron energy loss spectrometer. Setting Setting Cathode Monochromator Filament, V 3.22 Ml 2.639 Filament, A 1.90 M l -slit -0.31 Beam, E -6.09 Ml-cover -3.56 Emission Optics AMI-cover 0.16 R -2.7 1 M2 0.638 A1 46.22 M2-slit 0. 12 AA] -3.54 M2-cover -0. 16 A2 0.70 AM2-cover 0.03 AA2 l .60 M2-exit 0.49 A3 0.10 AA3 0.56 Lenses Analyzers B1 0.78 AN 1 0.572 AB 1 0.04 AN l-slit 0.46 B2 3.05 AN 1 -cover -0.43 B3 2.98 AANl-cover 0.14 B4 0.67 AN 2 0.576 AB4 0.1 1 AN 2-slit 0.35 AN 2-cover -0.43 Miscelaneous AANZ-cover -0.01 Shield 0.0 Sample 0.003 Ch-slit 1.00 Ch-cone 0.5 1 Scan 0.001 Scan-step 3.41 Offset 0.001 B3 ramp 1.0 B4 ramp 1.0 OS. ramp 1.0 141 Electron multiplier (Ch) Lenses (B l-B 2) Monochromator Pre-monochromator (M l) / Emission optics (R, Al-A3) Figure C.1 Schematic of the ELS3000 electron energy loss spectrometer. 142 llllllllllllllllllllllllzlllll 31293 02177