mmvyvv—u—u TU‘KV—‘— _—fi w—v qu ” ‘ .2' | u “m . -rm 1‘ l ‘ ‘A ‘h t . r I V A ' -‘ . ‘ . _ ' . z ' . . ‘ v. ‘ V . ’ , A ' ' . V ' ‘ A. ' ' ‘ ‘ h‘ ' H . f h . ‘ .Z ‘ ‘ > I - ‘v V ' I V ‘ ‘ . ‘ C". ' _ . . r , , , , ‘ . . . : v . . . t. , - ‘ , ‘ r V *. . - , , _ ‘t , ‘ ‘ ‘ . 4 _ ‘ _ ‘ 7‘ _: . _. ’ " ' : . , ‘ '. I -* < " ' z ‘ A , . ; . . \ ‘ ‘ ‘ . ‘ —. . a . , ' - - . ~ 1 '. . ‘ - .' A ’ , j ’ ‘ ‘ - . . .7 V ‘ . , ‘ \ ~ 7 . ' . i ‘ V .'. _ - A , - . ,- x . . . ‘ . .. , _,. ‘ A A A I . I V ‘ ‘ . ’ h < ' r 1 . . ‘ ~ ' ‘ " V . ' ' " 1 ' ‘ I ' 2 ‘ ‘ A ' ' ‘ > ! ‘ 7' I ‘ . ’ ~ ‘ "' u ’ ' , ‘ ‘ .‘v‘ .' . . , ‘ . . A - _ . - . . , ., . _ . . L 7 r. . 7 ‘. . s:.; h I H .‘ . , , >‘ ‘ ;>‘ ‘ . In .1 v . . A . ‘. , ‘ . . . . . . ‘ .'. -. ‘ ' ' . ‘1 * - " ‘ ‘u f _ ‘ A - ‘ . . ‘ ~ .v- ‘ n . .. 4‘ ,V‘ . . . ‘ , , . ;. . v ; . , . ‘ _ , , .7... .....‘. .. — org—3411‘ 5-54 ..‘ 1,“. . . - . - . .‘u ..n '1 .‘v u ‘ nu . . . ‘»;..‘....1-a ' -».u .r-- ‘ ‘ ‘ u. n .V. ‘ . ...... ..y. 7 n .‘ - .o: ‘ 7 . ‘ VERSITY LIBRARIES LIBRARY |\llllll\llllllllllllll W. \\ \fll ll .. Michigan State University This is to certify that the dissertation entitled Electron Dynamics as Observed by Adsorbate—Induced Broadband Infrared Reflectance Change Measurements presented by i Dennis Eugene Kuhl has been accepted towards fulfillment of the requirements for Ph.D. degree in Physics em: Roger G. Tobin Major professor Date August 7, 1996 MS U i: 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 duo. DATE DUE DATE DUE DATE DUE MSU Is An Affirmative MIoNEquol Opportunity Institution l Wanna-m ———-—~—~r._ ELECTRON DYNAMICS AS OBSERVED BY ADSORBATE-INDUCED BROADBAND INFRARED REFLECTANCE CHANGE MEASUREMENTS By Dennis Eugene Kuhl A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physics and Astronomy and Center for Fundamental Materials Research 1996 ABSTRACT ELECTRON DYNAMICS As OBSERVED BY ADSORBATE-INDUCED BROADBAND INFRARED REFLECTANCE CHANGE MEASUREMENTS By Dennis Eugene Kuhl In some cases gas adsorption on clean metal surfaces induces a negative broadband change in the metal’s infrared (IR) reflectance AR/ R which is large compared to the adsorbates’ dielectric properties. A model has been proposed. which explains AR/ R in terms of diffuse conduction electron scattering by the adsorbates and relates AR/ R to a wide variety of physical phenomena including a resistivity change. In this study I experimentally test the model by measuring AR/ R as a function of frequency, temperature, and coverage for CO on Pt(l 11). This marks the first time the scattering model has been applied to a transition metal. A second set of experiments tests the predicted linear relationship between AR/ R and tAp , the resistivity change (times the sample’s thickness), by measuring both effects simultaneously for O and formate adsorption on an epitaxial Cu(100) thin film. The frequency dependence of AR/ R for CO on Pt(111) is at least qualitatively consistent with the theoretical prediction. The magnitude of AR/R at 2500-2800 cm'I is smaller at 90 K than at room temperature by about 30%. This result appears surprising, since the scattering model predicts that |AR| should be larger at the lower temperature. The clean surface reflectance R, however, is also larger at low temperature, with the result that a decrease in IAR/Rl at these frequencies is consistent with the theory. The magnitude of the reflectance change peaks at a coverage of about 0.33 ML, and decreases toward saturation. The slope of the curve at low coverage corresponds to a scattering cross section per CO on the order of 1 A2. The presence of a peak in the coverage dependence could be attributable either to partial ordering in the overlayer or to changes in CO's electronic structure. Simultaneous measurements of AR/ R and tAp confirm linearity for 0 adsorption on Cu(100), although there is a large discrepancy between the predicted and measured slopes. The linearity test for formate adsorption on an O predosed Cu surface was inconclusive. As the formate replaces the O on the Cu( 100) surface, AR/ R appears to return very close to zero. tAp , however does not return nearly as close to zero. This difference, along with the possible lack of a linear relationship for formate, could indicate that the scattering model does not apply to formate adsorption on Cu(100). ‘B.N.J. Persson and AI. Volokitin, Surf. Sci. 310 (1994) 314 . to my Savior iv ACKNOWLEDGMENTS I want to first thank my wife, Lesley, for her support and encouragement throughout this Ph.D. project. She has been a wellspring of faith and love, as well as an example of perseverance. Her patience during the many days I was away completing this dissertation is greatly appreciated. Knowing that her prayers and thoughts were always with me provided me with powerful incentive to complete this long-term project. I owe a great debt of gratitude to my parents, Dale and Shelva Kuhl, for the firm foundation they provided me. It was only through their sacrifices that I was able to attend college and pursue the career in science of which I’d always dreamed. Moreover, their love, faith, and prayers have given me an anchor to depend on all my life. While thanking family members for their contributions to my efforts, I would be remiss if I didn’t mention my brother and sister, Douglas Kuhl and Stephanie Kuhl Lowman. I can always count on Doug to provide a respite from academic matters, usually via a golf game. Furthermore, he is always willing to help however he can, from moving my things to letting me stay over at his house in Canton. Steph and her husband Walt deserve recognition especially for their willingness to provide me with a place to stay on numerous occasions during the process of writing this dissertation, along with occasional meals. It was always relaxing to go to their home at the end of a frustrating day of writing. Two church congregations have contributed to the pursuit of my goals over the years. I was raised in the congregation at Moreland Christian Church, and it was through them that my system of values and spiritual foundation was built. Upon coming to graduate school I was very fortunate to have the opportunity to participate in the congregation at University Christian Church. They helped me to grow spiritually during the pursuit of my Ph.D. Numerous graduate students and friends have played a role in these efforts. First my Science Theatre friends: Bill Abbett, Dave Bercik, Danielle Casavant, Joy Conrad, Jennifer Discenna, Erik Hendrickson, Gerd Kortemeyer, Jeff Kriessler, Rod Lambert, Normand Mousseau, and Steve Snyder. Several friends provided me with a place to stay during the writing process while I made my home out of town: Larry and An Heimann (and Alex and Mark!), Jeff Schubert (the most comfortable sofal), Jeff Kriessler (too bad the organ wasn’t finishedl), Carl and Patricia Hoff (and Scott!). This work was supported in part by the Petroleum Research Fund, The Center for Fundamental Materials Research, and the NSF under Grant No. DMR-9201077 and DMR-9400417 (MRSEC). Helpful communications with CT. Campbell and M. I-Iugenschmidt (through Campbell) contributed to the closer modeling in Chapter 3. R. Naik graciously shared her expertise in the technique of growing the thin films. I want to thank Prof. Don Jacobs of the College of Wooster for encouraging me to pursue graduate school. The example he set of a teacher-scholar continues to be an vi inspiration to me. Prof. Jerry Cowen of Michigan State University was also a valuable source of encouragement. My coworkers in our research group deserve special citation. The UHV and IR apparati were assembled by Chilhee Chung. Keng-Ching “Kathy” Lin patiently instructed me in the operation of the system, and pioneered a great deal of the work contained herein. Hong Wang was always a source of encouragement. Larry Voice contributed to the thin film work and shared my love of junk food. Evstatin “Ati” Krastev developed the evaporation system, characterized the thin films, and ably took over the experiments when I moved on. I have saved for last the individual who deserves the greatest thanks, my advisor Prof. Roger Tobin. Without Roger, my pursuit of a Ph.D. would have gone nowhere. His creativity, knowledge, and expertise made this project possible. He was always willing to field any type of question, and could always be counted on to help me out of a jam. The patience of which he was capable was particularly impressive when an expensive piece of equipment fell victim to my inexperience. In many ways he set for me a high standard toward which to strive as a graduate student, and in the years to come. vii TABLE OF CONTENTS Chapter 1. Introduction Section l-l. Motivation for Studying Adsorbate-Induced Broadband Reflectance Changes Section 1-2. Adsorbate-Induced Broadband IR Reflectance Change in Metals Section 1-3. Adsorbate-Induced Broadband AR/R and Resistivity Change Ap in Metals References Chapter 2. Apparatus and Experimental Technique Section 2-1. Ultrahigh Vacuum Techniques and Cleaning the Pt(111) Surface Section 2-2. Reflection-Absorption Infrared Spectroscopy Section 2-3. RAIRS Measurement Technique Section 2-4. Thin Film Growth Section 2-5. Adsorbate-Induced Resistivity Change Measurement References viii 18 26 29 29 35 39 44 50 53 Chapter 3. Capillary and Effusive Gas Dosers 56 Section 3-1. Theoretical basis for closer modeling 58 Section _3-2. Calculations 62 Section 3-3. Results and discussion 64 References 71 Chapter 4. Infrared Reflectance Change Induced by CO Adsorption on Pt(111) 72 Section 4-1. Frequency Dependence 74 Section 4-2. Temperature Dependence 78 Section 4-3. Coverage Dependence 82 Section 4-4. Conclusions 91 References 92 Chapter 5. Changes in the DC Resistivity and the Infrared Reflectance Induced by Oxygen and Forrnate Adsorption on Cu(100) 95 Section 5-1. 0 Adsorption on a Cu(100) Thin Film 96 Section 5-2. Forrnate Adsorption on an O-Predosed Cu(100) Thin Film 105 References 109 Chapter 6. Conclusions 111 References 115 ix LIST OF TABLES Table 1-1. Comparison of predicted and measured reflectance changes AR/ R by Lin, Tobin, Dumas, Hirschmugl, and Williams in Reference 17. The quantity na is the adsorbate coverage. The values of tAp come from the literature. Table 1-2. Comparison of predicted and measured values of the ratio AR/Ap by Hein and Schumacher, reproduced from Reference 22. Table 3-1. Comparison of circular effusive doser arrays with various numbers of holes. All designs except those with eight holes comprise a single center hole surrounded by a ring of equally spaced holes at d/D = 0.8. The eight-hole arrays have only the ring of holes, without a center hole. It can be seen that increasing the number of holes from 9 to 17 has little effect. The center hole improves the enhancement factor slightly; it improves the uniformity for em, = 60° and degrades it for 0",“, = 40°, but in both cases the effects are small. Table 4-1. Bulk material parameters for Pt. Frequencies in s'1 have been divided by 21m to convert to cm". Table 4-2. Scattering cross section per adsorbate for several adsorption systems. 23 64 73 90 LIST OF FIGURES Figure 1-1. C=O internal stretch measured by RAIRS for 0.41 ML CO adsorbed on Pt(111). This plot is the fractional difference of a scan of a CO covered surface and a scan of a clean surface. Figure 1-2. Broadband reflectance change induced by 0.25 ML 0 adsorbed on Cu(100). This is a synchrotron-based RAIRS study performed by Lin, Tobin, and Dumas, reproduced from Reference 18. The gaps in the data and the scatter at high frequencies are artifacts of the polyethylene windows on the UHV chamber. Figure 1-3. Slab model for near-surface conductivity. l is the elastic mean free path, as and 0'3 are the near-surface and bulk conductivities respectively. Figure 2-1. A model of our Pt(111) surface. From Reference 2 by KC. Lin. Figure 2-2. AES for a clean Pt surface. The vertical axis is the derivative of the number of electrons detected at a particular energy, with respect to energy. Figure 2-3. TPD spectrum from a clean Pt(111) surface following 02 dosing at 90 K. Figure 2-4. The layout of the optical path. Reproduced from Reference 6 by C. Chung. Figure 2-5. An example of a reflectance time scan. The raw data are shown in (a), while the data in (b) are divided by the initial value with a linear baseline removed. The background pressure during the CO dose was 8.0 x 10'10 Torr. xi 19 31 33 34 36 41 Figure 2-6. Dependence of C=O stretch peak frequency on coverage for atop CO on Pt(111). From Ref. 23 by J .S. Luo, R.G. Tobin, and UK. Lambert. 43 Figure 2-7. A typical resistivity time scan. The data are divided by the initial value with a background slope subtracted. Background pressure during the dose was 5x10'6 Torr. 51 Figure 3-1: Origin of the cos3(0) term in the calculation of flux per unit sample area. 60 Figure 3-2: Normalized flux per unit sample area versus the squared distance from the sample's center for various sample to closer distances, indicated by 9mm: = arctan(D/2€). d/D is the ratio of the doser's radius to the sample's radius. Shown are plots for a capillary array of d/D = 2.0 and 1.0, a centered circular nine hole effusive array of d/D = 1.0 and 0.5, and a single effusive doser. 66 Figure 3-3: Area-weighted fractional standard deviation 0' versus the fraction of emitted molecules directly impinging on the sample f,- for various sample to closer distances, indicated by Omar = arctan(D/2€). d/D is the ratio of the doser's radius to the sample's radius. The data shown are for a capillary array of d/D = 0.4, 0.6, 0.8, 1.0, 1.2, 1.4, 1.6, 1.8, 2.0; a centered circular nine hole effusive array ofd/D = 0.3, 0.5, 0.65, 0.8, 1.0, 1.35, 1.7, 2.0; and a single effusive emitter. 68 Figure 4-1. Fractional reflectance change AR/ R as a function of frequency, for 0.33 ML CO on Pt(111) at room temperature. The lines represent fits of the Persson-Volokitin model1 to the data, with the specularity parameter p as the only adjustable parameter. The dotted line represents the best fit using DC parameters, and the solid line the best fit using optical parameters (see Table 4-1). 76 Figure 4-2. Dependence of the reflectance on Z/ 8 predicted for adsorption on Pt by the Persson-Volokitin scattering theory, for the indicated values of (II/(o, . The vertical scale for (a) and (c) is arbitrary. The small arrows indicate the room temperature (RT) and 90 K values of l/ 5 for Pt. DC values of the bulk parameters have been used (see Table 4-1). (a) Absolute reflectance change AR. (b) Clean-surface reflectance R. (c) Fractional reflectance change AR/ R. 81 xii Figure 4-3. (a) Fractional reflectance change AR/ R as a function of coverage, for CO on Pt(111) at 315 K. Data are shown for two IR frequencies, 2500 and 2800 cm". (b) Work function change for CO on Pt(111), reproduced from References 34 (Norton, et al., triangles) and 35 (Ertl, et _aI., circles). 86 Figure 5-1. Time scans of (a) the reflected IR intensity (i.e. voltage output from a lock-in amplifier) at 2800 cm'1 and (b) the voltage drop across the sample (i.e. . voltage output from a lock-in amplifier) with an applied current. Both 50 L doses involved background pressures of 5 x 10'7 Torr for 100 s. 98 Figure 5-2. Time scans of (a) the broadband reflectance change AR/ R at 2800 cm'1 and (b) the fractional change in resistivity Ap/ p . Each curve has a linear drift removed. Both 50 L doses involved background pressures of 5 x 10'7 Torr for 100 s. 99 Figure 5-3. AR/ R versus tAp for 02 adsorption on a Cu(100) film. The inset is an enlargement of the lower right hand corner. The solid line is a linear fit to the data, which are taken from about the 200 s to 500 S region of Figures 5-2(a) and 5-2(b). 102 Figure 54. AR/ R versus tAp for formate adsorbtion on a Cu(100) film. The inset is an enlargement of the upper left hand corner. The data are taken from about the 500 s to 620 8 region of Figures 5-2(a) and 5-2(b). 107 xiii Chapter 1 Introduction l-l. Motivation for Studying Adsorbate-Induced Broadband Reflectance Changes The study of the physics and chemistry of solid surfaces and interfaces, referred to as “surface science” due to its interdisciplinary nature, is of great importance both 2 The multibillion-dollar fields of lubrication and technologically and scientifically." sensor technology inherently involve surface properties. An understanding of surface processes such as catalysis, corrosion, oxidation, and crystal growth is essential to many diverse industries. On the other hand, most scientific work geared toward developing a basic understanding of the properties of condensed matter, both theoretical and experimental, tends to focus on bulk properties. This is due in part to the intrinsic difficulty of dealing with surfaces. In the words of Wolfgang Pauli:2 “The surface was invented by the devil.” Although the physical laws that describe interactions at surfaces are essentially the same as those for the bulk, the broken symmetry in the direction normal to the surface and the different environment of atoms at the surface greatly complicate theoretical treatments. Nevertheless, those same features lead to interesting new phenomena. In addition to the theoretical constraints, experimental research on surfaces has been historically inhibited by the difficulty of obtaining, characterizing, and maintaining clean surfaces for study. 2 While the thermodynamics of surfaces were nearly completely described by Gibbs3 in 1877, the complexity of performing calculations and the inability to experimentally test theoretical predictions limited progress in surface science for the next 80-90 years. In the last 30 years, however, several technological advances have Spurred a renaissance in surface science research. The three most significant advances involve vacuum technology, electron diffraction, and high-speed computers.2 First, through advances related to the space program, it became possible to commercially develop ultra- high vacuum (UHV) chambers in which one could obtain and maintain clean surfaces through cleaving or cleaning procedures. Second, knowledge of the submonolayer chemical composition of surfaces could be obtained through electron spectroscopy. Third, the availability of high-speed computers permitted numerical solutions to surface science problems. One of the most heavily studied areas of surface science is the interaction of gas atoms or molecules with metal surfaces, called adsorption. Surface vibrational spectroscopy (SVS) has been employed with great success to probe the vibrational resonances within the adsorbed molecule or between the molecule (or atom) and the substrate (i.e. the metal surface). SVS includes such techniques as electron energy loss spectroscopy (EELS), surface enhanced Rarnan spectroscopy (SERS), thermal energy atom scattering (TEAS), and reflection-absorption infrared Spectroscopy (RAIRS), among others.‘1 Studies utilizing these techniques have yielded a great deal of information regarding the fascinating complexities of the coupling and energy transfer between the vibrational and electronic states of the adsorbate and those of the substrate. 0.01 . I . I fl I e r 0.00 -0.01 -0.02 g -0.03 -0.04 0.41 ML -0.05 ‘ ~ CO/Pt(l 1 1) I I -0.06 .. -0.07 1 41 I 1 4 I A l A 2050 2070 2090 2110 2130 2150 frequency (cm'l) Figure 1-1. C=O internal stretch measured by RAIRS for 0.41 ML CO adsorbed on Pt(111). This plot is the fractional difference of a scan of a CO covered surface and a scan of a clean surface. Figure 1-1 shows an example of a RAIRS measurement of the internal stretching mode of CO adsorbed on a Pt(1 1 1) surface. The coverage is 0.41 monolayer (ML) CO. First a scan is conducted through the spectral region with the Pt surface clean. A second scan is conducted through the same spectral region after the surface is dosed with CO. In Figure 1-1 we represent the data in terms of the fractional reflectance change AR/ R by subtracting the “dosed scan” from the “clean scan,” and dividing by the “clean scan.” 4 Information about numerous physical processes can be gained by measuring quantities such as the peak frequency, the peak width, and the peak intensity while varying specific parameters, forexample the temperature or coverage. While using RAIRS to study such vibrational modes provides information about the coupling and energy transfer between the adsorbate and the substrate, using the technique to measure broadband changes in the reflectance allows the exploration of issues related to electron dynamics in the near-surface region of the substrate. In order to motivate the connection between the adsorbate-induced broadband change in a metal’s reflectance and electron dynamics, it is useful to discuss the current understanding of the effect of adsorbates on electron transport in metals. Even without strong energetic coupling adsorbates can profoundly affect electron dynamics in the near-surface region of a metal. An isolated adsorbate on an otherwise perfect surface breaks the translational symmetry parallel to the surface and sets up a static impurity potential from which conduction electrons can scatter without conserving the parallel component of momentum. In other words, the presence of an adsorbate changes the surface scattering of conduction electrons from specular to diffuse. These scattering events increase the electrical resistivity near the surface, with observable and sometimes dramatic consequences. It has been known for many years that even a monolayer of adsorbed gas can increase by more than 20% the resistance of a metal film hundreds of atomic layers thick.5'7 Indeed, measurement of this physical property forms the basis for an important class of chemical sensors. It was recognized by Holstein8 and othersg'” as long ago as 1948 that diffuse scattering of conduction electrons from surfaces would necessarily also 5 reduce the infrared reflectance of the surface. Only recently, however, have such broadband changes in the IR reflectance been measurable. 2 5 '3 0 I o . have revrved Interest In Recent experiments12 '24 and theoretical developments these scattering effects as they relate to adsorbate-substrate dynamics. A number of careful infrared studies on single crystals in UHV have revealed adsorbate-induced decreases (typically ~l%) in the broadband reflectance AR/ R of metal surfaces. The dielectric properties of the adsorbate layer itself would be expected to produce an increase in the reflectance, but at least two orders of magnitude smaller than the observed effects.31 The observed reflectance changes, therefore, result from changes in the electronic properties of the substrate caused by the binding of the adsorbates. Figure 1-2 is an example of one of the recent measurements of a broadband change in the IR reflectance. It is reproduced from Reference 18, a synchrotron-based study of 0 adsorption on Cu(100) by Lin, Tobin, and Dumas. Figure 1-2 shows the data for 0.25 ML 0 on a Cu(100) surface. The data is collected by first performing a scan with a clean surface, followed by a scan with a dosed surface. As in Figure 1-1, Figure 1- 2 shows the fractional reflectance change AR/ R by subtracting the dosed scan from the clean scan and dividing by the clean scan. The gaps in the data and the scatter at high frequencies result from the use of polyethylene windows on the UHV chamber, which have their highest transmittance in the low-frequency regime. Note the characteristic shape of the frequency dependence. Also note that the size of the effect, ~ 1%, is considerably smaller than the peak in Figure 1-1. 0.000 _ - -0.003 .. :, - g -0.006 ..'.. “23' "if? . 3. 51:31.5! : . -0.009 Tr-" ._ s - to..,, . "a: 2‘: .‘1 -0.012 - I— 9 = 0.25 ML -0015 + 3 4 O 400 800 1200 1600 2000 Frequency (cm-1 ) Figure 1-2. Broadband reflectance change induced by 0.25 ML 0 adsorbed on Cu(100). This is a synchrotron-based RAIRS study performed by Lin, Tobin, and Dumas, reproduced from Reference 18. The gaps in the data and the scatter at high frequencies are artifacts of the polyethylene windows on the UHV chamber. It is worth remarking in passing that measurements of broadband changes in surface reflectance are far more demanding experimentally than the more common observations of relatively sharp vibrational bands. The broadband measurements demand that the absolute intensity be held constant (or independently measured) to an accuracy considerably better than 1% between the measurement on the clean surface and that on the adsorbate-covered surface. Such stability is not easily achieved. Vibrational measurements, on the other hand, are relatively insensitive to changes in signal level so 7 long as the background remains relatively smooth. Additionally, investigation of the frequency dependence of the broadband AR/ R requires the ability to probe very low frequencies. Few RAIRS experiments have successfully yielded information below 1000 cm". Persson and Volokitinzs'29 have recently proposed a model based on these ideas of conduction electron scattering which accounts for the adsorbate-induced change in the broadband IR reflectance, AR/ R , of metals. Using the model, they develop a relationship between AR/ R and the adsorbate-induced change in resistivity, Ap. Furthermore, within the model they relate26’28'29 the adsorbate-induced AR/ R and Ap to the observation of antiabsorption resonances with RAIRS, the damping of hindered vibrational modes, and atomic scale friction. While it remains possible that electron scattering does account for all these effects, it has been shown17 that the elements of the model which relate AR/ R and Ap do not rely on the connections to the other phenomena treated. Regardless of the validity of the connections to the other effects, experimentally testing the predicted relationship between AR/ R and Ap , along with predictions regarding just the broadband AR/ R , can help establish whether or not scattering is the dominant mechanism. This knowledge will greatly contribute to our understanding of the interactions of gas molecules and atoms with metal surfaces. 1 have used RAIRS to perform a series of measurements of the broadband AR/ R which test the validity of the Persson-Volokitin scattering model. In Chapter 4, frequency-, temperature-, and coverage-dependent measurements of AR/ R for CO adsorption on a Pt(1 1 1) single crystal are presented and the results are compared to the 8 model’s predictions. In Chapter 5, simultaneous measurements of the adsorbate-induced AR/ R and Ap for O and formate adsorption on an in situ grown epitaxial Cu(100) film are presented. That data is used to test the predicted relationship between AR/ R and Ap. The remainder of Chapter 1 introduces the relevant parts of the Persson-Volokitin scattering model. The expression for AR/ R which comes out of the model is presented in Section 1-2. Recent measurements of AR/ R to which the model has been applied are discussed. In Section 1-3 the expression relating AR/ R and Ap is developed, along with a discussion of attempts to experimentally test it. 1-2. Adsorbate-Induced Broadband IR Reflectance Change in Metals The primary motivation for the development of the Persson-Volokitin model”29 of conduction electron scattering from adsorbates comes from several remarkable observations of rather strong antiabsorption resonances associated with hindered rotational and translational modes of adsorbates on metal surfaces. Antiabsorption resonances are spectral features which exhibit an apparent increase of the reflectance on resonance relative to nearby frequencies, as opposed to the usual decrease due to the absorption of light by a resonant mode. In 1988 Reutt, Chabal, and Christrnanl4 reported measurements of formally dipole-forbidden low-frequency frustrated translations of H on W000) and H on Mo(100). In 1990 Hirschmugl, Williams, Hoffrnann, and Chabal” reported measurements of the dipole-forbidden frustrated rotation of CO on Cu(100) at low-frequency. It is noteworthy that both experiments revealed a broadband absorption of IR light. 9 The dipole selection rule for IR detection of vibrational modes comes from the fact that the component of the incident IR electric field parallel to the surface, E", is strongly screened inside the metal. Since E” is continuous at the surface, the electric field on the vacuum Side is almost entirely perpendicular to the surface. Therefore, only modes with a dynamic dipole moment normal to the surface are observable with RAIRS.32 Persson’s idea regarding the observation of dipole-forbidden modes in the form of antiabsorption resonances was that even though E” is strongly screened inside the metal, its finite penetration can induce an oscillating drift current. The carriers of this drifi current scatter diffusively from the adsorbates, causing an increase in the surface resistivity of the metal. A consequence of the resistivity increase is a decrease in the metal’s broadband reflectance. When the frequency of the incident IR light matches the frequency of a hindered parallel mode of the adsorbate, however, the molecules move in resonance with the drift motion of the electrons. Thus the additional surface resistivity vanishes and the IR reflectance returns to its clean surface value. This model has done a good job accounting for the measurement of the frustrated rotation of CO on Cu(100) an Cu(11 1).19 The elements of the model which are tested here are the predictions regarding the broadband decrease in the IR reflectance. The Persson-Volokitin scattering model has its origins in the work of Fuchs,33 10,11 . on the anomalous skin effect. The Holstein} Reuter and Sondheimer,9 and Dingle anomalous skin effect is a surface property concerning electron transport and requiring nonlocal corrections to the classical Fresnel formulae. The classical Fresnel formulae, which depend on the bulk dielectric function e , often provide an adequate description of 10 the reflectance of a metal’s surface using a local form for e . A local description is valid when the condition 0) >> y—F— 1-1 8 is met, where vF is the Fermi velocity (i.e. the maximum velocity of a conduction electron), and 6 is the classical skin depth. Equation 1-1 requires that the incident E- field does not vary appreciably over the distance an electron travels during one cycle of the incident field. The distance an electron travels in the bulk between elastic scattering events is given by the elastic mean free path l, which takes the form I = th B , 1-2 where r B is the bulk scattering time. For most metals, 5 is on the order of hundreds of A and l is at least comparable to 6. In some cases, such as Cu, 6 is much larger than 8 . For low frequencies, where the condition given in Equation 1-1 is not met, a local description of e , and consequently the reflectance, is no longer valid. This is because an electron traveling into the near-surface region will experience a force from the penetrating E-field and propagate ballistically into the bulk. This is called the anomalous skin effect and the resulting transfer of energy from the incident light to the metal causes the metal’s ll reflectance to be less than unity. Moreover, it accounts for an intrinsic frequency- dependence in the reflectance. In addition to making nonlocal corrections to the Fresnel formulae, Holstein studied the general case where some fraction of the conduction electrons, given by the quantity (l-p), scatter diffusively from surface imperfections. The parameter p is know as the Fuchs specularity parameter and represents the probability of specular scattering. Persson adapted this treatment, interpreting adsorbates as surface imperfections. The scattering model of Persson and Volokitin incorporates a number of Simplifying assumptions: 1. The metal is treated as a free electron (Drude) gas described by an electron density n and a bulk scattering time 13, or, equivalently, a plasma frequency 03p and a mean- free-path l. 2. The adsorbate affects only the scattering rate, represented for the most part simply through a change in the Fuchs specularity parameter33 p, which represents the probability of specular scattering from the surface; (1 — p) is the probability of diffuse scattering. 3. The adsorbates form a dilute, disordered layer, so that scattering from the individual adsorbates is independent. The scattering probability can then also be described by a scattering cross section per adsorbate, 2: 12 (1- p) = n02 1-3 where no is the surface density of adsorbates. With these assumptions the fractional reflectance change AR/R for p-polarized light can be written:28 (£)=__3_VL(1-__’L).f[m g] 1.4 R 4c c050 (ll—1’5 where 0 is the angle of incidence of the light. Note that V; within a Drude model depends only on n, the conduction electron density. The frequency dependence is a universal function28 f that depends on the frequency 0) relative to 001 and the ratio of the electron mean-free-path l to the classical skin depth 8 = c/mp. (01 is a characteristic rolloff frequency given by v 0) (0' = F p , 1-5 c Near 001 an electron travels a distance comparable to the skin depth during one cycle of the electromagnetic field; (01 therefore specifies the frequency range below which nonlocal electrodynarnic effects are important.”28 13 At sufficiently high frequencies f is of order unity; in the approximation that the clean surface reflectance is one it approaches unity exactly, but this is not always a good assumption. It, is important to note that the parameters that enter into f do not involve properties of the adsorbate; the frequency dependence depends only on the metal, while only the magnitude of the reflectance change depends on the adsorbate. It is also worth noting that Equation 1-4 does not require the Specific model of adsorbate vibrational damping advanced in Reference 29. The complete analytical expression for AR/ R is / ‘l a... (myKellie/0.46)] .6 R — 2 i , - ( 1:00 p c030 8(0/(0195/5ML where g(0)/001, 13/5) and 8((0/031, lit/54) are complex-valued functionsz8 given by g=jdy(1-}17)F2(y)+mg 4mm Dd), , y -0)2-i0)11 [yi_%) 170)]: 1-7 and e=q2-££(£) 23+[[-i1]-1]1n .3 , 1-8 4B0, “I B I3 'q where 14 o l l F(y)=idq8(m/wra€/5rq)l+ifiy 1-9 and -——-. 1-10 The parameter a is the magnitude of the reflectance change at high frequencies, which is given by the prefactor of f in Equation 1-4. a = 3vF(1- p) 4ccosO In addition to predicting the frequency-dependent broadband AR/ R, Equations 1-7 to 1-11 also relate AR/ R to the damping of hindered translations and rotations 24,30,34 (observed as antiabsorption resonances). The first term in the expression for g(c0/0)1,€/5) in Equation 1-7 essentially contains the information about AR , while the second term gives the antiabsorption feature at 00 o. The parameter 1] is the damping coefficient associated with the resonance. The parameter 0) 0 does not come into play when calculating just the broadband AR/ R. Whether or not n is essential to calculating the broadband AR/ R remains a controversial point.”37 As will be discussed in Chapter 4, it is possible to set 11 equal to zero in the above expressions and still accurately predict the frequency-dependence of AR/ R. While that would seem to be conclusive evidence that n is not intrinsically related to AR/ R, Persson claims that the quantity (l-p) is 15 proportional to n .25 Nevertheless, the model has been very successful in explaining the 16,19 existence and line Shapes of the antiabsorption resonances and in accounting for the spectral Shape of the broadband reflectance.’8"9 The earliest reports on adsorbate-induced broadband reflectance changes dealt with adsorbates on W(100), by Riffe, Hanssen and Sieversn’l3 and Reutt, Chabal and Christrnan. '4 Electron scattering effects were considered, but the spectral Shape, which showed an increase in the reflectance change (back toward zero) at high frequencies, led Reutt et al. to interpret the results in terms of surface electronic states. Persson’s scattering model was applied in the analysis of data collected by Hirschmugl et al. for CO on Cu(100) and Cu(111).’6"9 It was found to do a good job fitting the frequency dependence in both cases. Another stringent test of the frequency dependence was performed by Lin, Tobin, and Dumas.l8 Figure 1-2 shows one of their measurements of the broadband reflectance change, together with a fit to the Persson-Volokitin model.27 The agreement with the predicted functional form is excellent, and the values of 0), are largely, though not completely, in agreement both among experiments and between experiment and theory.18 The model was unable to fit their data, however, for their highest coverage, 0.35 ML. Lamont, Persson, and Williams38 measured the broadband AR/R caused by H adsorbed on Cu(111). They found that scattering model works for low frequencies, but AR/ R increases much more rapidly for 0) > 1000 cm'l than predicted by the model. Following the example of Reutt et al.,'4 they tentatively interpret this behavior in terms of adsorbate-induced surface states. 16 Thus far I have only discussed the frequency dependence predicted by the scattering model. Because Equation 1-8 contains the ratio [/6 , and the elastic mean free path I is a strongly temperature-dependent quantity, there is a natural temperature dependence in AR/ R. Plots of AR , the absolute change in reflectance, which appeared in Reference 28 and were parametrized by the ratio €/5 would lead one to believe that for decreasing temperature (i.e. increasing [/8 ), AR/ R , the experimentally measured quantity, should get larger. The magnitudes of both AR and R increase as the temperature decreases, however, and a proper consideration of the clean surface reflectance R shows that the behavior of AR/ R depends on the particular temperature and frequency regimes in which the measurements are taken. The temperature dependence will be further discussed in Chapter 4, Section 4-2. The scattering model also predicts a linear dependence of AR/ R on coverage. This can be seen by combining Equations 1-3 and 1-8 to get a: ”FE no. [1-12] 4cc030 Recall that a is the magnitude of AR/ R at high frequencies and na is the number density of adsorbates. In their work, which predated the Persson-Volokitin scattering model, Riffe, 12,1 Hanssen, and Sievers 3 reported a coverage dependent adsorbate-induced broadband absorption for N2, 02, CO, H2, and D2 on W(100) using measurements of the 17 attenuation of surface electromagnetic waves. With increasing coverage, the magnitude of the attenuation coefiicient increased monotonically for N2, while a peak was observed for both 02 and CO. H2, and D2 adsorption each showed a slightly more complicated, nonmonotonic dependence on coverage. Pointing to earlier theoretical work,"'l "33 Riffe et al. attributed the broadband absorption primarily to surface-reconstruction-induced changes in the scattering of conduction electrons. They found correlations between the structure they observed in the coverage-dependent absorption and ordered patterns deduced from low energy electron diffraction (LEED) experiments.39 Using these correlations, they explained the observed structure in terms of ordering in the adlayer. For CO on Cu(100) and Cu(l 1 l), Hirschmugl et 01.15.16 observed the predicted linear dependence up to saturation. For 0 on Cu(100), however, Lin, Tobin, and Dumasl8 saw a roughly 20 % decrease, well outside experimental error, in AR/ R for their highest coverage. It is interesting that this was also the coverage at which the scattering model failed to fit the frequency dependence. In the present work I extend the investigation of the Persson-Volokitin scattering model to a transition metal, Pt, and explore the dependence of the reflectance change AR/ R on frequency, temperature, and adsorbate coverage. Since it is not clear that the Drude model should work for transition metals, this set Of experiments should help to establish the bounds of usefulness of the Persson-Volokitin scattering model. 18 1-3. Adsorbate—Induced Broadband AR/R and Resistivity Change Ap in Metals Following the example of Lin, Tobin, Dumas, Hirschmugl, and Williams,17 below I develop the relationship between the adsorbate-induced fractional change in the reflectance AR/ R and the adsorbate-induced change in resistivity Ap. The starting point is to expand the appropriate classical Fresnel formula40 for p-polarized light using the Feibelman d-parameter formalism.“ =__ , 1-13 c 0050 95 40) Im(d") R where 0) is the frequency of the light, c is the speed of light, and 0 is the angle of incidence. Equation 1-13 holds when the magnitude of the complex dielectric function e of the metal meets the condition |e| >> (1/ cos2 0)>> 1. The parameter (1" is given by 1-14 where o u is the complex conductivity parallel to the surface for a uniform electric field, and z is the direction normal to the surface. The Im(dl) is considered to be negligible compared to Im(d“) due to the strong refraction by the metal which makes E i very 19 small just inside the metal. From Equation 1-13 it can be seen that the problem reduces to calculating o ll(z). The effect of adsorbates on the electrical conductivity is felt in the near-surface region to a depth on the order of the elastic mean free path I, defined in Equation 1-2. Further inside the metal, the conductivity takes the bulk value 0'3. Persson proposes a slab model29 as illustrated in Figure 1-3. Figure 1-3. Slab model for near-surface conductivity. 1’ is the elastic mean free path, 0‘s and GB are the near-surface and bulk conductivities respectively. This model provides the following form for the surface conductivity to a depth on the order of the mean free path 3: o,(m)=oB(0))+Ao(m). l-15 20 This slab model provides a useful means of mapping a thin film, with which it is straightforward to measure the adsorbate-induced resistivity (conductivity) change, onto the near-surface region of a bulk crystal. Using Equation 1-15 for the conductivity to a depth 2 and GB elsewhere, d” can be calculated from Equation 1-14 and inserted into Equation 1-13 to get ARE = 4m(t/c)lm[Ac(co)]. 1-16 If the frequency is restricted to a region where the classical skin depth 6 >> vF /0) (i.e. the same condition as in Equation 1-1), a local description of the metal’s dieletric response can be used. This “high frequency” condition implies that an electron does not travel a significant distance during one cycle of the incident electric field. The Drude model for the bulk conductivity is a useful approximation to describe the local response in many metals: __n__8213 _in8), 1-17 03(9)): m( where n is the conduction electron density and m is the electron mass. Expanding A0 to first order in the changes An and Ar gives 21 A0 ___(__(0)_ An At 70(0)) n8 —+ ”(l—ions) 1-18 Since the scattering model assumes that the dominant process in the conductivity change is scattering, the first term in Equation 1-18 can be neglected. Expressing things in terms of resistivities rather than conductivities, Equations 1-16 and 1-18 lead to W —— Here p(€) refers to the resistivity of a film of thickness Z and p3 refers to the bulk resistivity. It is useful to note, however, that the resistivity of a film of arbitrary thickness can be used because the quantity tAp(t) is independent of thickness 1. This property can be seen in the Fuchs-Sondheimer model for thin film conductivity as well as through 5,7 experimental evidence. This, along with Equation 1-16, allows the fractional reflectance change to be written E = _ 4" AP“) = -————.[tAp(t)] 1-20 R c: BeosG p 8 —,,cp 1:0089 The only term in Equation 1-20 that refers to a thin film property rather than a bulk property is tAp(t). This is made possible through the adoption of a slab model. 22 Furthermore Equation 1-20 predicts the relationship between AR/ R and tAp(t) to be independent of frequency. Making use of p Blr = '£2 a 1'21 ne we get AR 4ne2 J __ = _ tA . 1-22 R (me c080 [ p] Note that this derivation avoids any assumptions regarding the damping of vibrations. The first attempt to test this relationship was performed by Lin et al.1.] They measured the broadband AR/ R for single coverages of CO and O on a Cu(100) single crystal, and for CO on a Ni(100) single crystal, and compared their results to values of AR/ R predicted by Equation 1-22. The predicted values were calculated from literature values of tAp , measured on polycrystalline thin film samples. The results of Lin et al. from Reference 17 are presented in Table 1-1. As can be seen in Table 1-1, they found reasonable agreement for CO and O on Cu(100), but a discrepancy for CO on Ni(100). There are inherent difficulties with the comparison, however. Lin et al. point out that there exiSts a difference in crystallographic orientation; the films are polycrystalline but believed to be primarily of (111) orientation, while the bulk single crystals were oriented in the (100) direction. 23 Additionally, variations in sample preparation and uncertainties in surface structure and sample thickness make the values of tAp reliable to only within about a factor of two, as indicated by the. values of tAp for O on Cu listed on Table 1-1. Table 1-1. Comparison of predicted and measured reflectance changes AR/ R by Lin, Tobin, Dumas, Hirschmugl, and Williams in Reference 17. The quantity na is the adsorbate coverage. The values of tAp come fiom the literature. "a (AP (AR/R)ca]c (AR/ R)8Xp system (1014 cm’z) (10'12 0 cm2) (%) (%) CO/Cu(100) 6.1 0.60 -2.2 -1 . 1:1:0.2 -1 .1101" O/Cu(100) 3.8 0.42 -1.5 -1.1i0.2 0.70 -2.6 CO/Ni(100) 6.4 1.47 -5.8 -0.2i0.2 . . . . 21 21corrected for a difference in ineldence angle M. Hein and D. Schumacher recently sought to alleviate the problem of using measurements on different samples by making simultaneous measurements on in situ grown thin film samples.22 They studied 0, CO, and Cu on Cu films, and Ag on Ag films. From their data they were able to plot the reflectance change versus the resistivity change to test the predicted linear relationship. They found strong evidence of linearity between tAp and AR/ R for O and CO on Cu, although they found the slope to differ 24 significantly from the predicted value. For Cu on Cu and Ag on Ag, they observed the resistivity reach a maximum at some particular coverage, afier which it decreased. They found that the linear relationship holds for coverages below the resistivity maximum. In the case of Ag on Ag they found a large discrepancy between the predicted and measured values of the slope. The results of the slope comparison of Hein and Schumacher from Reference 42 are presented in Table 1-2. Table 1-2. Comparison of predicted and measured values of the ratio AR/Ap by Hein and Schumacher, reproduced from Reference 22. (AR/AP)... (AR/49).... system (149 ' cm)” (119 ' cm)_l O/Cu -0.03O -0.0085 CO/Cu -0.014 -0.0077 Cu/Cu -0.010 -0.0085 Ag/Ag -0.025 -0.0082 In Chapter 5 I will present a test of the linearity prediction for O and formate adsorption on a Cu( 100) thin film. This experiment differs from the Lin, et al. '7 study and the Hein and Schumacher study22 in that the Cu(100) film is grown epitaxially on a 25 Si(100) substrate. Therefore this is the first simultaneous test of the predicted linear relationship between AR/ R and tAp on a well characterized thin film sample. 26 References lG.A. Somorjai, Chemistry in Two Dimensions: Surfaces (Cornell University Press, Ithaca, New York, 1981). 2 A. Zangwill, Physics at Surfaces (Cambridge University Press, Cambridge, 1988). 31 .W. Gibbs, Collected Works, Vol. 1 (Longmans-Green, New York, 1928) 184. 4J.W. Gadzuk, in Vibrational Spectroscopy of Molecules on Surfaces, ed. J .T. Yates, Jr. and TE. Madet (Plenum, New York, 1987). 5 P. Wissmann in: Surface Physics, ed. G. Htihler, Springer Tracts in Modern Physics Vol. 77 (Springer, New York, 1975). 6D. Dayal, H.-U. Finzel and P. Wissmann in: Thin Metal Films and Gas Chemisorption, ed. P. Wissmann (Elsevier, Amsterdam, 1987). 7D. Schumacher, Surface Scattering Experiments with Conduction Electrons, ed. G. Hdhler, Springer Tracts in Modern Physics Vol 128 (Springer, New York, 1993). 8T. Holstein, Phys. Rev. 88 (1952) 1427. 9G.E.H. Reuter and EH. Sondheimer, Proc. Roy. Soc. London, Ser. A 195 (1948) 336. ”RE. Dingle, Physica 19 (1953) 311. ”RB. Dingle, Physica 19 (1953) 729. l2D.M. Riffe, L.M. Hanssen and A.J. Sievers, Phys. Rev. B 34 (1986) 692. l3D.M. Riffe, L.M. Hanssen and A.J. Sievers, Surf. Sci. 176 (1986) 679. ”1.13. Reutt, Y.J. Chabal and SB. Christrnan, Phys. Rev. B 38 (1988) 3112. 15OJ. Hirschmugl, G.P. Williams, F.M. Hoffmann and Y. J. Chabal, Phys. Rev. Lett. 65 (1990) 480. “’CJ. Hirschmugl, Y.J. Chabal, F.M. Hoffrnann and G.P. Williams, J. Vac. Sci. Technol. A 12 (1994) 2229. 27 17KC. Lin, R.G. Tobin, P. Dumas, C.J. Hirschmugl and GP. Williams, Phys. Rev. B 48 (1993)2791. "’K.C. Lin, R.G. Tobin and P. Dumas, Phys. Rev. B 49 (1994) 17 273 (1994); ibid. 50 (1994)17760. I9C.J. Hirschmugl, G.P. Williams, B.N.J. Persson and A. I. Volokitin, Surf. Sci. 317 (1994) L1141. 2"DM. Riffe and A.J. Sievers, Surf. Sci. 210 (1989) L215. 2’13. Borguet, J. Dvorak and H.L. Dai, SPIE Proceedings 2125 (1994) 12. 22M. Hein and D. Schumacher, J. Phys. D: Appl. Phys. 28 (1995) 1937. 23C.L.A. Lamont, B.N.J. Persson and G.P. Williams, Chem. Phys. Lett. 243 (1995) 429. 24J. Krim, D.H. Solina and R. Chiarello, Phys. Rev. Lett. 66 (1991) 181. 25B.N.J. Persson, Chem. Phys. Lett. 197 (1992) 7. 26B.N.J. Persson, Chem. Phys. Lett. 185 (1991) 292. 27B.N.J. Persson and AI. Volokitin, J. Electron Spectrosc. Relat. Phenom. 64/65 (1993) 23. 28B.N.J. Persson and AI. Volokitin, Surf. Sci. 310 (1994) 314 . 29B.N.J. Persson, Phys. Rev. B 44 (1991) 3277. 301.8. Sokoloff, Phys. Rev. B 52 (1995) 5318. 3|R.G. Tobin, Phys. Rev. B 45 (1992) 12 110. ”Yr. Chabal, Surf. Sci. Repts. 8 (1988) 211. 33x. Fuchs, Proc. Cambridge Phil. Soc. 34 (1938) 100. 34B.N.J. Persson, 1. Chem. Phys. 98 (1993) 1659. 35R.G. Tobin, Phys. Rev. B 48 (1993) 15 468 36B.N.J. Persson, Phys. Rev. B 48 (1993) 15 471 28 37T.A. Gerrner, J .C. Stephenson, E.J. Heilweil and RR. Cavanagh, J. Chem. Phys. 101 (1994) 1704. 38CA. Lamont, B.N.J. Persson, and GP. Williams, Chem. Phys. Lett. 243 (1995) 429. 39DA. King and G. Thomas, Surf. Sci. 92 (1980) 201. 40W.L. Schaich and W. Chen, Phys. Rev. B 39, (1989) 10714. "Pr. Feibelman, Prog. Surf. Sci. 12 (1982) 287. Chapter 2 Apparatus and Experimental Technique When used to study the interactions between adsorbed gases and surfaces, reflection-absorption infrared spectroscopy (RAIRS) is typically performed on ordered surfaces of bulk Single crystal samples. This helps to make analysis of the results tractable; single crystal surfaces facilitate the application of theoretical results to physical systems. The production of clean surfaces suitable for study on such samples requires intensive cleaning procedures and ultrahigh vacuum (UHV) conditions. A second option for the production of suitably clean, ordered surfaces is to grow epitaxial thin metal films in situ in UHV. Both large single crystals and thin metal films have been used in this study. Section 2-1 describes the UHV apparatus, the Pt(1 11) single crystal sample, and the surface cleaning techniques. Section 2-2 tells about the home-built RAIRS apparatus while section 2-3 details the RAIRS measurement technique. Cu(100) thin film growth is discussed in section 2-4. Finally, section 2-5 describes the resistivity change measurement. 2-1. Ultrahigh Vacuum Techniques and Cleaning the Pt(lll) Surface If each impinging molecule sticks to the sample surface (i.e. sticking coefficient is one) and the surface atomic density is about 3 x 10”cm '2 (both of which are reasonable physical assumptions in many cases, depending upon the mass of the gas, among other 29 30 variables), an exposure of 1x10'6Torr-s, defined as a Langmuir, will result in a monolayer coverage on the sample surface. Thus, a clean metal surface at a pressure of 1x 10"°Torr will require a little more than two and one half hours to build up a monolayer of contamination. Our experiments were typically done at a base pressure of about 4x 10'1 1Torr. The stainless steel UHV chamber is 12” (30.5 cm) in diameter and has a height of 25” (98.4 cm). It is rough pumped by a molecular sorption pump and a turbomolecular pump. During normal UHV operation the chamber is pumped by ion pumps1 and by a titanium sublimation pump (as needed). A bakeout at about 100°C for 36-48 hours is required to achieve UHV. This is accomplished by means of a series of heater tapes wrapped around the chamber, an internal infrared (IR) lamp, and a heater around the ion pumps. During bakeout the chamber is wrapped in aluminum foil to more evenly distribute the heat and the temperature is monitored by an array of thermocouples. From bottom to top the chamber is divided into the IR/evaporator level, the surface analysis level, and the manipulator. The IIUevaporator level will be discussed separately. The surface analysis level includes instrumentation for ion sputtering, low energy electron diffraction (LEED), and Auger electron Spectroscopy (AES). There is also a residual gas analyzer which enables temperature programmed desorption (TPD). The manipulator allows the sample to translate in the X, Y, and Z directions as well as rotate in the azimuthal angle 4). The sarnple used in the single crystal experiments was a 1 cm x 3 cm Pt single crystal with a surface normal 27° from the [111] direction, as determined by Laue X-ray 31 diffraction and LEED. The lower right comer was cut away because it exhibited a greater degree of miscut from the [111] direction. The remaining 2.7° miscut results in a roughly In other words, there is a one-atom step about every twenty 5% density of _step sites. 2 surface Pt atoms, as shown in Figure 2-1. Figure 2-1. A model of our Pt(1 1 1) surface. From Reference 2 by KC. Lin. We do not expect the ~ 5% density of step sites to affect our results because the step sites should be saturated with CO at the coverages studied. Moreover, the change in scattering rate due to CO on the terrace should be independent of the presence of scattering centers on the step edges (Mathiessen's rule). For purposes of this study, then, the sample can be regarded as a Pt(1 l 1) crystal. 32 The sample was mounted at the end of a stainless steel tube, which could act as a liquid nitrogen cold finger, on the manipulator. The sample was heated radiatively by a tungsten filament and the temperature was measured by two chromel-alurnel thermocouples spot-welded to the edge of the sample. Temperatures from ~ 90 K to ~ 1400 K were routinely accessible. The sample could be dosed with various gases by means of a leak valve and a multihole effusive doser array,3 which provided a flux enhancement of about 16 over background dosing. Details of doser design are given in Chapter 3. AES was used to determine the contaminants present on the surface and thus the necessary cleaning protocol. For example, large amounts of Ca, which segregated from the bulk, required cycles of Ar+ sputtering and armealing to 1175 K. This typically left a great deal of C on the surface. Exposure to 02 at various elevated temperatures was an effective way to remove the C, but often left a small amount of O on the surface which Was difficult to remove. Surface 0 can usually be removed by heating to 1360 K. In cases where this was not effective, the O was probably bonded to another surface contaminant such as Si, which is difficult to observe with AES because the primary Si peaks overlap with strong Pt peaks.4 In these instances it was necessary to sputter the surface again and repeat the entire procedure. When the AES indicated no contaminants above the noise level (~ 4% of the Pt 237 eV peak) the surface was considered clean. Figure 2-2 shows an ABS scan that meets this criterion. Note that the positions of the peaks associated with the major contaminants are indicated in the figure. 33 0.18 . . a - . . , . , 0.14 - - 0.10 - 0 j C Ca j ‘ g 0.06 r- l l m i 1 E L z 0.02 - 'O . T 1 T . -0.02 - j Pt 1 i 1 Pt Pt - . Pt“ Pt j Pt . -0.06 - - . 1 t P _010 l 1 Pt I t. l r l r l 100 200 300 400 500 electron energy (eV) Figure 2-2. AES for a clean Pt surface. The vertical axis is the derivative of the number of electrons detected at a particular energy, with respect to energy. Unfortunately, the hot filament used in AES itself tended to deposit traces of C on the surface. Therefore, the last sequence performed consisted of exposure to oxygen at 90 K followed by TPD and flash heating to 1360 K. A secondary indication of a clean surface was when repeated 90 K 02 closing and TPD showed that the 02 peak had reached a maximum. A typical TPD spectrum following 90 K 02 dosing of a clean Pt(lll) surface is shown in Figure 2-3. Note that the 02 peak dominates all other monitored masses. H2 and a gas with a mass of 28 amu, either CO or N2, each Show small peaks at 34 low temperature. The fact that the peak occurs at low temperature indicates that it probably corresponds to N2. I I V l I 2.0 - - 1.6 - . ”a? 3 <— 02 '3 1.2 r - E9 5 H20 g 0.8 L . L CO 01' N2 0 4 - n. H2 - t‘ ‘4 ’W 0.0 a l a l a l m 0 200 400 600 800 1 000 temperature (K) Figure 2-3. TPD spectrum from a clean Pt(1 11) surface following 02 dosing at 90 K. As stated above, an ordered surface with its normal in the (111) direction was indicated for our Pt sample by its hexagonal LEED pattern. Doublet spots confirmed the roughly 5% miscut from the (111) direction calculated from Laue X-ray observations. 35 We were also able to observe two ordered patterns for CO adsorbed to the Pt(1 l 1) surface which have been previously studied.5 At a coverage of 0 = 0.33 ML a diffuse («5 x J3)R30 pattern was observed. A fairly sharp c(4 x 2) pattern was observed at a coverage of 0 = 0.5 ML. 2-2. Reflection-Absorption Infrared Spectroscopy Figure 2-4, which is reproduced from Reference 6 by C. Chung, shows the layout of the spectroscopic system. The instrument is a home-built apparatus with a spectral range of ~ 300 - 3000 cm", and is designed for use in the study of low-frequency vibrations of atoms and molecules at surfaces. While many design choices were made with that specific mission in mind, the instrument was successfirlly utilized in this study to measure the broadband change in reflectance, as will be described in Section 2-3. A conventional silicon-carbide globar heated resistively to about 1300 K serves as the IR source. The globar is situated at one focal point of an externally adjustable gold- coated ellipsoidal mirror. The IR light from the source is intensity-modulated with a tuning fork chopper7 which is placed at the other focal point of the ellipsoidal mirror. The tuning fork chopper requires no lubrication and very low power, making it ideal for operation in vacuum at low temperature. It chops the light at a frequency of 800 Hz, which is chosen to avoid mechanical resonances in the instrumentation. The source housing, chopper, and mirror are rigidly mounted to a copper plate which is in contact with a liquid nitrogen reservoir. The plate is supported by three rods kinematically mounted to the bottom of the source cryostat. The source housing, chopper, and mirror 36 are cooled to liquid-nitrogen temperature in high vacuum to reduce modulated radiation from sources other than the globar as well as atmospheric absorption of the light. FAR- IWRARED SLRFACE MCIROSCOPY SYSiEM Liquid NlUO‘lICII l.t)‘"Cd Spectral rmge: .11!) ~ .1000 aI-l (stating petlfmltflt'l Resolution: 1-5 ‘CI'I SWIG 1w: 20 ~ 1200 K -._. "1;. ;_—JV Ultrmiql Vaculn Surface Analysis Elmer uith Single Crystal Staple on Lite-cooled manipulator 1".1' '«I.11,'tl l .!1 l1.ll 11 vi: Delel lur (\ .3 E35 ’Inr' V-— 1:; \. ‘4’“ . .. i R C t t Other surface proves. . .__. _, d ' 509’“? W05 0, low inetqy Electron Uillnlt't i011 _ Ill" lN-coaled 5’1'9'95- Auger Electron ‘upet‘llanopy f- ‘0 CI, ”4 i’OlUf l/tfl (lfld 010111le “1911111” [#330er 1011 'lpr’iil OSLOpy Figure 2-4. The layout of the optical path. Reproduced from Reference 6 by C. Chung. Within the same high vacuum chamber, the IR light passes through the chopper to the transfer optics. These consist of two gold-coated, off-axis parabaloidal mirrors, the first of which can be adjusted from outside the vacuum. The transfer optics magnify the image of the chopper by a factor of 1.8 so that the IR light will interact with as many 37 adsorbates on the sample surface as possible. The light is incident on the sample at 86° from the surface normal through a differentially pumped CSI window8 on the UHV chamber. CSI is chosen due to its high transmittance down to low frequencies. The reflected IR light passes out of the UHV chamber through a second CSI window into a complementary transfer optics which are a mirror image of the first pair. Thus they demagnify the IR light by a factor 1.8 and focus it onto the entrance slit of the spectrometer. The Spectrometer, the heart of the apparatus, is a home-built Czemy-Turner grating spectrometer housed in high vacuum and cooled to liquid nitrogen temperature. As with the source, the vacuum serves to reduce atmospheric absorption as well as provide thermal insulation for the cold components. Cooling the spectrometer reduces the background photon flux that reaches the detector. The optical components are rigidly attached to a large aluminum plate which is in contact with a liquid nitrogen reservoir. The optical plate sits on three invar rods kinematically mounted to the bottom of the spectrometer shell. During normal operation the optical plate is clamped to the outer shell for mechanical stability via three steel screws. A variety of entrance slits for the spectrometer are mounted on an adjustable wheel at the focus of the transfer optics, allowing the resolution of the spectrometer to be changed externally. After the entrance slit the light encounters a gold-coated folding mirror which directs it toward the collimating mirror. The collimating mirror is a gold- coated, off-axis parabaloidal mirror, focused at the entrance slit. The now collimated light next diffracts from the grating, which is a gold-coated replica on an aluminum 38 substrate.9 Three different gratings are used to cover the entire spectral range and the system must be opened to change gratings. The diffracted light is collected and focused by the camera mirror, which is also a gold-coated, off-axis parabaloid. Following the camera mirror is another folding mirror which directs the light through a polarizer. The polarizer consists of a gold wire grid with a line spacing of 2880 lines/mm on a silver bromide substrate.10 Finally, the p-polarized component of the light focuses on the exit slit wheel. Similar to the entrance slit wheel, the exit slit wheel is externally adjustable. The exit slit wheel, however, includes a variety of low-pass IR filters which enable the attenuation of higher orders of diffraction. Upon leaving the spectrometer, the light enters a final transfer optics consisting of a gold-coated, externally adjustable, ellipsoidal mirror. The exit slit of the spectrometer is situated at one focal point of the ellipsoidal mirror, and the detector is situated at the other. The transfer optics is in the same high vacuum Space as the spectrometer, the UHV charnber-to-spectrometer transfer optics, and the detector, and has its own liquid nitrogen reservoir. Just before the detector in the optical path is a Winston cone. A Winston cone is a parabolic cone designed such that any light entering the wide end within the angular acceptance of the cone undergoes successive reflections and emerges from the narrow end. It collects the IR light and delivers it to the small area of the detector element without forming an image. The Winston cone has a gold-coated interior surface, an entrance aperture of 10 m, an exit aperture of 1.0 mm diameter, and a focal ratio of 4.4. The IR detector is an extrinsic Si:B photoconductor operated at liquid He temperature.11 The signal fi'om the detector is amplified by a two-stage transimpedence 39 amplifier (TIA) with two switchable feedback resistors. The high and low feedback resistors are mounted on the liquid He temperature cold plate to reduce Johnson noise and can be chosen via a latching relay. The first stage of the amplifier uses a matched pair of J230 JFETS mounted on the liquid He cold plate and heated to 77 K. The second stage is a conventional AC coupled operational amplifier at room temperature with a gain which is switchable between two and ten. Section 2-3. RAIRS Measurement Technique The detector is normally operated at a DC bias of 2.50 Volts. When the spectrometer is warm, or when the grating is at zero degrees, the detector must be set to low feedback. High feedback can be used when the spectrometer is cold because the background photon flux is so much smaller. The chopped signal is taken to a lock-in amplifier12 and demodulated using the chopper output as the reference. The IR intensity data are collected from the lock-in amplifier by computer as a function of either frequency (grating angle) or time (for a fixed grating position). A vibrational measurement involves scanning through the spectral region of interest once when the sample surface is clean, dosing with a gas, and scanning through the same spectral region a second time. Absorption peaks or other spectral features due to adsorption of the gas are found when the data are represented as a fractional difference (i.e. the clean scan is subtracted from the scan of the closed surface, and the difference is divided by the clean scan). It is possible to choose a particular frequency of interest and monitor the intensity of the reflected light as a function of time while dosing with the gas. This type of scan is referred to as a “time scan.” If a frequency away from sharp spectral 40 features is chosen, one can use a time scan to observe in real-time the change in intensity of the broadband reflectance (at a single frequency) caused by the adsorption of a gas on a metal surface. Both spectral scans and time scans were used in this study. A typical data taking procedure for a broadband measurement on the Pt(1 1 1) sample went as follows. First the sample was flashed to 975 K to clear the surface of a passivating layer of CO. The monochromator was set to a single IR frequency and the IR intensity was monitored for 5-30 minutes, during which interval the sample was dosed with CO. Finally, TPD to 975 K was performed, which provided information about the coverage of CO during the measurement and prepared the sample for the next measurement. Figure 2-5 shows the change in the reflected IR intensity versus time for a single measurement. Figure 2-5(a) is the raw data, showing the voltage output of the lock-in amplifier. In Figure 2-5(b) the data are normalized to the initial value, with a systematic linear baseline drift subtracted. During this run the sample was closed with CO for 100 seconds beginning at about the 200th second at a background pressure of 8.0 x 10"0 Torr. The pressure at sample was enhanced by about a factor of about 16 due to the closer (see Chapter 3). The fractional change in reflected IR intensity for this run, AI / I , and therefore the fractional reflectance change, AR/ R, was 0.00200 i 0.00002. 41 0.03526 . r . r, ~ . r (a) * 0.03524 - . 3 0.03522 . ~ s. '2 .5 l i 8 1 fig 0.03520 - 1 1 - a: 1 3 l0.42i:0.03 ML CO/Pt(l 11) l _8 0.03518 - 2800 cm-1 i ‘13 315 x 1 1 1 0.03516 - q dosing ‘- 0.03514 . ' . L. . 1' ~ ' . 0.0005 * I ' f; ' I I r I f l ! ‘1 .‘ 1 ’ ‘ ll 11 l I a” 0.0000 1' i . 1 ‘.Id ”H" ”i. ' Ill 1 l H 11 1 ~ l 11'] g A 1 . . l * -0.0005 - ' ; Ti ‘ 1 i l : ‘ S -0.0010 - 1 ; 0.200%10.002% - <1 . 1 1 04210.03 ML CO/Pt(1 11) l ? -0.0015 - ' . - 2800 cm1 i z . . , : 4 315 x 'L _0.0020 b .1, Ill Alli“ I.‘ l i' i Ii 111411 l l 1111 -0.0025 . . 3 fl dosrng A‘— 0 100 200 300 400 500 time (s) Figure 2-5. An example of a reflectance time scan. The raw data are shown in (a), while the data in (b) are divided by the initial value with a linear baseline removed. The background pressure during the CO dose was 8.0 x 10'[0 Torr. 42 This particular run exhibited an unusually stable baseline, which can be observed in Figure 2-5(a) . In most cases baseline drifts of 0.2% to 2% over the course of a run limited the accuracy. We attribute the majority of this drift to a progressive loosening of the electrical contacts on the globar as it is cycled through temperature extremes in normal use. Tightening the contacts tended to reduce the observed drifi, but doing so was inconvenient because it required disassembly of the source cryostat. Even larger drift was observed when the coldfinger was filled with liquid nitrogen due to sample drift resulting from expansion and contraction of the coldfinger as the liquid nitrogen boiled away. Usually the drift from both origins was quite linear which allowed us to subtract the baseline from the measurement, and the reflectance change could be measured to an accuracy of 0.02% or better. It is of interest to compare the sensitivity of our measurements to that of the U4IR beamline at the National Synchrotron Light Source (N SLS) of Brookhaven National Laboratory, where most of the previous broadband reflectance studies have been carried 13-18 out Although the IR flux from the synchrotron is much greater than in our system, electron beam movement and uncertainties in beam current limit the sensitivity to broadband changes to about 0.2%,”’18 nearly an order of magnitude above our level. The changes observed here for CO on Pt(1 l 1), which have a maximum magnitude less than 0.3%, would be extremely difficult to measure with the synchrotron system. A significant drawback of our system, however, is that we measure only a single IR frequency in each run, whereas the FTIR spectrometer at the NSLS obtains the entire spectrum at once. 43 The CO coverage was determined with reference to the C=O stretching frequency of atop-bonded CO, an example of which is shown in Chapter 1, Figure 1-1. This absorption peak appears around 2100 cm'l and is rather sharp (i.e. a width of S 10 cm"). It has been well studied (References 19-23 provide just a small sampling of the literature concerning the C=O stretch mode), so the dependence of the peak frequency on coverage is well known. When measuring AR/ R via time scans at 2500 cm“1 and 2800 cm'l, we first measured the clean-surface reflectance through the spectral region 2050 to 2150 cm'l. After the time scan, during which CO was dosed onto the sample, we again scanned the same region to record the atop-bonded C=O stretching vibration. ‘The saturation coverage at room temperature was assumed”25 to be 0.5 ML and the rate of change of frequency with coverage was taken from Figure 2-6, which comes from Reference 23 and is a compilation of published data. 2110L- r + ' _. 2‘ +01: 's -: c u q o e e V t. v E "’ o C, + a O a +" a. x g 53° 9.. x 11.2090—" x x — x to x or (L x x 2080 L 1 0 0.3 0.6 Total CO Coverage (ML) Figure 2-6. Dependence of C=O stretch peak frequency on coverage for atop CO on Pt(1 l 1). From Ref. 23 by J.S. Luo, R.G. Tobin, and D.K. Lambert. Measurements at lower frequencies required different gratings, and the C=O stretch region could not be measured; for these measurements a dosage vs. coverage curve determined from the higher-frequency measurements, as well as integrated TPD signals, were used to determine the coverage. Uncertainty in the actual room temperature saturation coverage of our sample could introduce an overall shift in the coverage scale. Broadband reflectance change measurements on the thin fihn samples followed a very similar protocol. There were, however, additional considerations in timing the operations which resulted from the steps needed to grow a new film sample for each measurement and to make a simultaneous resistance measurement. 2-4. Thin Film Growth As shown in Chapter 1, the scattering model predicts a linear relationship between the adsorbate-induced IR reflectance change, which is a bulk quantity, and the adsorbate- induced resistivity change, which is typically measured on thin film samples. In order to examine this relationship experimentally, it is necessary to choose a sample which is thick enough to allow reflectance measurements to probe the bulk reflectance, yet thin enough for the change in resistance due to the presence of adsorbates to be measurable. The relevant length scale for reflectance measurements is the classical penetration depth 5 , which for Cu at room temperature15 is 270 A. The sample should be thicker than this to satisfy the requirement that it be “bulk-like.” The scale of the resistivity change is related to the elastic mean free path; the adsorbate-induced fractional change in resistivity 45 decreases in proportion to the ratio of If to the thickness,26 l/ t. The elastic mean free path for Cu at room temperature is about 1140 A. We have chosen to grow Cu thin film samples about 500 A thick on hydrogen-terminated silicon surfaces prepared by wet chemical etching, which have been shown to be excellent substrates for growing epitaxial Cu(100) films.2’°3" E.T. Krastev made a number of modifications to the UHV chamber to equip it for thin film growth.38 A sample transfer system was added which includes a turbomolecular-pumped loadlock, a magnetically coupled linear translator, and a custom- designed home-built sample holder. This allows the rapid transfer of new Si substrates into the UHV chamber. The Si substrate is mounted on a small disk which is inserted into the UHV chamber through the loadlock via the linear translator. Three spring-loaded posts and a keyhole locking system affix it to the sample holder. This sample holder replaces the one used in the single crystal experiments, but attaches to the end of the same liquid nitrogen coldfinger and manipulator. The sample substrate has Six electrical contacts: two for a thermocouple and four for a four-probe resistance measurement. As in the single crystal experiments, the sample is heated radiatively by a tungsten filament. The minimum temperature is limited by the quality of thermal contact allowed by the spring tension holding the sample disk in place. The maximum temperature is limited by the tendency of the springs to degrade at high temperatures. Temperatures of about 110 K to 775 K can be regularly employed. An evaporator was added38 which includes two evaporation filaments, a rotating shutter, a crystal thickness monitor, and appropriate shielding. The thickness monitor and 46 shielding can be cooled by flowing water or liquid nitrogen. Recently, Krastev again modified the evaporator to include a loadlock.38 It is now possible to remove the evaporator filaments to reload the evaporation metal without breaking the UHV. In preparation for a measurement, the (100) Si substrates (B-doped, 20—50 S) cm resistivity) are cleaved to the appropriate size for the sample holder and then prepared for insertion into the vacuum chamber. The Si is then degreased in an ultrasonic cleaner using first acetone and then methanol. Before etching, electrical contacts for the four- probe resistance measurements are deposited onto the Si(100) substrates in a high vacuum (HV) evaporation chamber. Using an aluminum foil mask wrapped around the Si, 100 A of Cr followed by 1500 A Ag are deposited onto the edges of the Si. The Cu films are deposited so that they overlap the contacts on the ends. Crucial steps in the preparation of the Si are etching for 30 - 60 seconds in a 10% aqueous solution of HF, and pull-drying (slowly and smoothly removing the substrate from the solution with the surface vertical, so that the liquid sheets off smoothly with no droplets). The etching and pull-drying lead to an extremely flat and chemically inert surface, with virtually all the Si dangling bonds terminated”42 with H. After etching, the samples are loaded into the loadlock and pumping is begun as quickly as possible. After transferring the Si substrate into the UHV chamber, the substrate is heated to 700 K for 10 minutes to outgas it. AES has shown that this leads to the production of the cleanest Cu surface; the largest remaining contaminants tend to be C and O at a level of less than 4% of the Cu 60 eV LMM AES peak. When fihns are grown without first outgassing the Si, the C and 0 levels can be as high as 30% - 40% of the Cu 60 eV LMM ABS peak. In order to sustain the outgassing temperature without overheating the 47 electrical feedthroughs or springs, it is necessary to cool the coldfinger with liquid nitrogen. The liquid nitrogen also helps bring the sample’s temperature down after the outgassing. Since the best film growth is achieved at room temperature,36 the sample temperature is chosen to be about room temperature by balancing the heat lost to the coldfinger with heat added by the sample heater. Following the outgassing, it normally takes about one hour to stabilize the temperature. Once the temperature is stabilized, the film can be deposited. A Significant amount of work has been done by E.T. Krastev and L.D. Voice to characterize the crystal structure, surface morphology, and electrical conductivity of these films as a function of deposition rate, deposition temperature, and post-deposition annealing.36 Some of the results will-be discussed here. The best films we grow are obtained near room temperature at growth rates of 1.0-10 A /s. Higher temperatures and higher deposition rates produce lower quality films. Post-deposition annealing, while expected to improve the smoothness of the surface, can only be done at relatively low temperatures due to the formation of copper silicides at the Si-Cu interface at higher temperatures. Temperatures up to 125°C have little effect, so post-deposition annealing is not used in this work. While my interest is in producing high quality films for use in reflectance/resistance studies, the Cu(100) on Si(100) growth system is unusual and interesting in its own right because there is a 40% mismatch between the two lattices. This fact makes the relatively easily achieved epitaxial growth of Cu(100) on Si(100) curious. The existence of a mixed "buffer layer" about 100 A thick between the Si and the Cu was discovered by an earlier study.43 This buffer layer is quite likely important to 48 the epitaxial growth. Reflection high-energy electron diffraction43 (RHEED) and grazing incidence X-ray diffraction”44 studies have shown that the Cu lattice is rotated by 45° relative to the Si lattice, with Cu(010) parallel to Si(011). The rotation reduces the lattice mismatch between the two materials from 40% to 6%. Even a 6% mismatch, however, is unusually large for epitaxial growth. The strain is probably at least partially relieved by the buffer layer. For most of the characterization work, E.T. Krastev deposited the copper films in a resistive evaporator with a HV base pressure 1 X 10"8 Torr, although later films were prepared in the 10'll Torr UHV chamber. The UHV films show test results indicating that they are no worse than the high vacuum films in terms of structure and morphology, and are usually among the best. Thus, the high vacuum film results can be considered essentially valid for the UHV films as well. The film orientation in the direction perpendicular to the surface plane was characterized by standard 0-20 X-ray diffraction (XRD). Comparing36 XRD scans of completely disordered (i.e. powdered) bulk Cu, polycrystalline Cu films, and Cu films grown on oxidized (i.e. unetched) Si substrates to the scans of films grown on etched Si substrates demonstrated highly oriented growth in the direction perpendicular to the surface, with the Cu(100) direction aligned with Si(100). 45-48 Four The inplane orientation was determined by means of X-ray pole figures. very distinct poles were observed which confirm that the films posses very highly oriented cubic structure and therefore they are epitaxial. Quantitatively, more than 97% of the copper is oriented.38 49 For the UHV films we were able to examine the surface order qualitatively using low energy electron diffraction (LEED). At room temperature we successfully observed the square pattern expected for a Cu(100) surface. The surface morphology was studied in air using a commercial atomic force microscope49 (AF M). The primary measure of surface roughness used was the root- mean-square (RMS) deviation of the height from the mean. Scans of bare Si substrates yielded RMS roughness values of 2.0 - 3.0 A. Typical AFM images for Cu samples grown at deposition rates of 1.0 -2.0 A /s at room temperature exhibit a granular structure with RMS roughness values of 10 - 20 A. Although the films are epitaxial, they are relatively rough on the atomic scale. Finally, L.D. Voice performed four-wire resistance measurements on selected films during HV deposition, which allowed the fihn resistance to be continuously monitored as a function of film thickness. He used the growth conditions which had been demonstrated to produce the best films in terms of crystal structure and surface morphology.36 The results indicated a high degree of scattering, either from surface defects or internal grain boundaries, was likely. Nevertheless, for the highest thicknesses the conductivity approaches a value of 55 (u!) - m)-l , which agrees with the value of o for pure Cu at the deposition temperature, 54 (u!) - m)". 50 2-5. Adsorbate-Induced Resistivity Change Measurement Once the new film is deposited, it is necessary to wait for the UHV chamber to attain an acceptable base pressure, and for the sample to stabilize at the desired temperature. The pressure typically reaches 1x10'10 Torr after approximately 20 minutes, at which point the measurement is started. During the waiting period the sample must be moved to the IR position and the spectrometer must be prepared for a measurement by setting the frequency and maximizing the light intensity that reaches the detector. When all systems are ready, simultaneous reflectance and resistivity time scans are begun. The internal reference of a lock-in amplifier12 (LIA), set to 1.00 V at 1.0 KHz or 3.0 KHz, is applied to the sample. This provides approximately 1.7 mA current to the sample due to the 600 (2 output impedance of the LIA which is large compared to the sample’s resistance, typically 2 - 5 Q. The voltage drop across the sample is measured by the LIA using two shielded coax cables, with the shields grounded, for leads. The LIA reads the difference between the two leads, which helps to eliminate noise common to both. The capacitive and inductive pickup in the leads were both measured to be about 8x10'9 V/J Hz , which is only slightly above the limit of sensitivity of the LIA, 5x10'9 V/JH—z at 1 KHz. The data is collected from the LIA by a computer using a program similar to the one used for reflectance time scans. As with the reflectance time scans, for resistivity time scans it is important to measure the clean-surface value long enough to establish a stable baseline. The sample is then dosed with the desired amount of gas by backfilling the UHV chamber. The doser is not used in the thin fihn experiments due to Space constraints in the UHV chamber; with 51 the closer in place it is impossible to transfer a sample between the loadlock and the sample mount. Following the dose, the resistance is monitored long enough to again establish a stable baseline in both time scans. Figure 2-7 is an example of a resistivity time scan. 0.12 ' I ‘ I ' I ‘ I ' I ' I ' I 0.08 - . i 1 g 0.06 P _, dosing (<— - 4 500 L CO dose 004 _ 530 A Cu(100) - 300 K 0.02 - - 0.00 ~=— - 0 100 200 300 400 500 600 700 800 time (s) Figure 2-7. A typical resistivity time scan. The data are divided by the initial value with a background slope subtracted. Background pressure during the dose was 5x10'6 Torr. 52 In Figure 2-7, the data are divided by the initial value and a linear baseline is subtracted. The 500 L CO dose was achieved by backfilling to a pressure of 5x10’6 Torr for 100 5. During the dose Ap/ p quickly reaches a plateau. After the dose ends a surprising thing happens in this run: Ap / p actually decreases. This could result from an oversaturation of CO on the surface such that the coverage decreases due to CO leaving the surface as the CO pressure in the UHV chamber decreases. Other possibilities include relaxation of the CO overlayer or adsorbate-induced reconstruction of the Cu surface. Simultaneous measurements of reflectance and resistivity will be discussed in more detail in Chapter 5. ' 53 References lPerkin-Elmer Corp. TNBX-1000 ion pump system. 2RC. Lin, Ph.D. thesis, Michigan State University, 1995, unpublished. 3D.l~:. Kuhl and R.G. Tobin, Rev. Sci. Instr. 66 (1995) 3016. 4LE. Davis, N.C. MacDonald, P.W. Palmberg, G.E. Riach, and RE. Weber, Handbook of Auger Electron Spectroscopy, 2nd ed. (Physical Electronics Industries, Inc., Eden Prairie, MN, 1976). 5H. Steininger, S. Lehwald and H. Ibach, Surf. Sci. 123 (1982) 264. 6C. Chung, Ph.D. thesis, Michigan State University, 1993, unpublished. 7Multi-Scanning Systems Corp., type RC-2. 8R.G. Tobin, C. Chung and J .S. Luo, J. Vac. Sci. Technol. A 12 (1994) 264. 9Milton Roy Co. loPerkin-Elrner Corp. 11Infrared Laboratories, Inc. '2 Princeton Applied ResearchCorp. l3C.J. Hirschmugl, G.P. Williams, F.M. Hoffinann and Y. J. Chabal, Phys. Rev. Lett. 65 (1990) 480. “or. Hirschmugl, Y.J. Chabal, F.M. Hoffrnann and GP. Williams, J. Vac. Sci. Technol. A 12 (1994) 2229. ”KC. Lin, Ra. Tobin, P. Dumas, C.J. Hirschmugl and GP. Williams, Phys. Rev. B 48 (1993) 2791. 16KC. Lin, R.G. Tobin and P. Dumas, Phys. Rev. B 49 (1994) 17 273 (1994); ibid. 50 (1994) 17760. 54 17Cl Hirschmugl, G.P. Williams, B.N.J. Persson and A. I. Volokitin, Surf. Sci. 317(1994)L1141. 18C.L.A. Lamont, B.N.J. Persson and GP. Williams, Chem. Phys. Lett. 243 (1995) 429. , 198.13. Hayden and AM. Bradshaw, Surf. Sci. 125 (1983) 787. 20M. Tushaus, E. Schweizer, P. Hollins, and AM. Bradshaw, J. Electron Spect. 44 (1987) 305. 21CW. Olsen and RI. Masel, Surf. Sci. 201 (1988) 444. 221D. Beckerle, R.R. Cavanaugh, M.P. Casassa, E.J.Heilweil, and J.C. Stephenson, J. Chem. Phys. 95 (1991) 5403. 231s. Luo, R.G. Tobin and D.K. Lambert, Chem. Phys. Lett. 204 (1993) 445. 2“1.). Malik and M. Trenary, Surf. Sci. 214, (1989) L237. 25L.F. Sutcu, J.L. Wragg and l-I.W. White, Phys. Rev. B 41 (1990) 8164. 26P. Wissmann, in Surface Physics, edited by G. Hahler, Springer Tracts in Modern Physics 77 (Springer, New York, 1975). 27CA Chang , Phys. Rev. B 42 (1990) 11 946. 28CA Chang, Surf. Sci. 237 (1990) L421. 29C..A. Chang, J. Appl. Phys. 68 (1990) 5893. 3"CA Chang, J. Vac. Sci. Technol. A 8 (1990) 3779. 3’C.- A. Chang, J. Vac. Sci. Technol. A 9 (1991) 98. 32Y.-T. Cheng, Y.-L. Chen, M.M. Karmarkar and W.-J. Meng, Appl. Phys. Lett. 59 (1991) 953. 33Y.-L. Chen and Y.-T. Cheng, Mat. Lett. 15 (1992) 192. 34Y.-T. Cheng and Y.-L. Chen, Appl. Phys. Lett. 60 (1992) 1951. 35R.Naik, M. Ahmad, G.L.Duifer, C.Kota, A.Poli, Ke Fang, U. Rao and J.S. Payson, J. Magnetism and Magnetic Mat. 121 (1993) 60. 55 36E.T. Krastev, L.D. Voice, and R.G. Tobin, J. Appl. Phys. 79 (1996) 6865. 37Minsu Longiaru, R.G. Tobin, E.T. Krastev, and L.D. Voice, J. Vac. Sci. Technol., to be published. 38E.T. Krastev, Ph.D. thesis, Michigan State University, in preparation. 39T. Takahagi, I. Nagai, I. Ishiytani, H. Kuroda and Y. Nagasawa, J. Appl. Phys. 64 (1988)3516. 40T. Takahagi, I. Ishiytani, H. Kuroda, Y. Nagasawa, H.1to and S.Wakao, J.Appl. Phys. 68 (1990) 2187. “YJ. Chabal, G.S. Higashi, K. Raghavachari and VA. Burrows, J. Vac. Sci. Technol. A 7(1989)2104. 42P. Dumas, Y.J. Chabal and GS. Higashi, Phys. Rev. Lett. 65 (1990) 1124. 43BCDemczylr, R.Naik, o. Auner, C.Kota and U.Rao, J. Appl. Phys 75 (1994) 1956. 44CA Chang, J.C. Liu, and J. Angilello, Appl. Phys. Lett. 57 (1990) 2239. 45J.S.Kallend, U.F.Kocks, A.D.Rollett, H.-R.Wenk, Materials Science and Engineering, A132 (1992) 1. 4(U.F.Kocks, J.S.Kallend, H.-R.Wenk, A.D.Rollett, S.I.Wright, popLA manual, Los Alarnos National Laboratory, July 1994. 47H.-R.Wenk, Preferred Orientation in Deformed Metals and Rocks: An Introduction to Modern Texture Analysis, Academic Press, New York, 1985. 48M Hatherley, W B Hutchinson An Introduction to Texture in Metals, The Institution of Metallurgists, Monograph No 5. 49Digital Instruments Nanoscope III Scanning Probe Microscope. Chapter 3 Capillary and Effusive Gas Dosers In many surface science experiments, including those that comprise this study, it is necessary to expose a surface to a known and uniform flux of gas. Often dosing can be accomplished simply by filling the vacuum chamber to a uniform pressure. When dealing with highly reactive gases, gases that are not pumped effectively, gases that exchange rapidly on the chamber walls, and very high flux levels, however, it is necessary to use a doser that provides a high flux at the sample while keeping the background pressure in the chamber low. An example where a doser is needed which is relevant to this study is 02, which is important in the Pt surface cleaning process and has a low sticking coefficient on Pt. Dosers are typically one of two types: effusive, in which the aperture diameter is much larger than its length, or capillary, in which the diameter is much smaller than the length. The effusive doser is usually a single effusive aperture, whereas the capillary doser is commonly used either singularly (the “needle” doser) or in an array. The process of designing and implementing a closer appropriate for our UHV system led us to investigate the relatively unexplored possibility of arrays of effusive dosers. The primary criteria in designing a doser are the uniformity of the flux on the sample and the enhancement factor. The enhancement factor is defined as the ratio of the flux at the sample to the flux on a surface elsewhere in the chamber, or equivalently the 56 57 ratio of the "effective pressure" at the sample to the rise in the background pressure. The design of a doser inevitably involves a tradeoff between these two criteria. In previous work, Campbell and Valonel published a theoretical analysis comparing capillary arrays with a single effusive source. They calculated flux distributions and enhancement factors for circular capillary arrays ranging from a single capillary to an array with a diameter equal to the sample diameter, for a range of sample- to-doser distances. Their results Show that a single needle closer2 gives a very large enhancement factor but an extremely nonuniform flux distribution. They also concluded that a single effusive source (“cosine emitter”) is in many cases as good as a capillary array in both uniformity and enhancement factor. To obtain reasonable uniformity, a capillary array needs to be at least as large as the sample; no results were given, however, for arrays larger than the sample. Winkler and Yates4 considered larger arrays but calculated only the enhancement factor, not the flux distribution. Here we extend Campbell and Valone's calculations to larger capillary array sizes. In the following, all dimensions are assumed to be much smaller than the mean free path of the gas molecules. We is also show that a single effusive source is far less effective than their curves suggest, and indeed is usually inadequate unless a very large flux gradient across the sample is tolerable. We present flux calculations for effusive arrays consisting of a relatively small number of effusive sources situated near the perimeter of the sample, and Show that such arrays can be competitive with capillary arrays if the closer is not too close to the sample. This Chapter is based on work published in Reference 4. 58 3-1 Theoretical basis for doser modeling The starting point is the calculation of F(0), the integral-normalized flow of molecules per unit solid angle from a single source, at an angle 0 to the source axis: 9 3-1 where 1(0) is the flux in molecules-s"-sr", and Nm, is the total flow out of the source in molecules-s". For an effusive source F (0) is F(9)= €089 . 3-2 7! and for a single straight capillary of length L and radius a it is:5 nap; 2cos0 W 2 ; cos0 F(9)= “W {(1——)R(p)+$(l-W{l—Q—pZY]+ }09e nsrnO 2 59 where p = L(tan 0)/2a, R(p) = COS'1(p) — p(1-—p2)‘/2v 0,, = tan'1(20/L), and W = (8a/3L)(l + 8a/3L)'1. As other authors, such as Benziger and Madix, have noted,5 the flux from a capillary of large L/a is highly collimated. For example, for Ma = 40, 97% of the molecules emerge within a cone of half-angle BC = 29°. Little is gained by using larger L/a ratios; the central peak becomes sharper, but the fraction of flux at larger angles does not change. For this reason, we will follow Campbell and Valone in using L/a = 40 throughout this work. The highly collimated flow from a single capillary has been cited1 as demonstrating that a single needle doser is unsatisfactory. This conclusion is correct but probably overstated. The expressions for F(G) for 0 < BC in Equation 3-3 assume that some molecules can go directly through the capillary from the reservoir behind it, without hitting the walls.5 This is generally correct for capillary arrays, but a needle doser more ofien consists of a tube with at least one bend between the open end and the gas reservoir. Such a geometry will lead to a flux distribution less strongly peaked in the forward direction than that given by Equation 3-3. Nevertheless it will be more directional than a single effusive source, which we Show below is already unacceptably nonuniform for most applications. From a practical point of View the flux per unit solid angle is of less interest than the integral-normalized flux per unit area of the sample, which we denote G(0). Figure 3-1 illustrates the geometry of a planar sample oriented perpendicular to the axis of the emitter. 60 sample Figure 3-1: Origin of the cos3(0) term in the calculation of flux per unit sample area. Using this geometry, it is easy to show that the integral-normalized flux per unit area of the sample is given by 0(a) = F(e)cos30 . 34 The extra factor of cos30 has a noticeable but modest effect on the rather directional flux from a capillary, but a severe effect on the flux from an effusive source. In fact, it is somewhat misleading to refer to the single effusive doser as a “cosine emitter” when one 61 considers the extra factor of c0530. The flux uniformity plots shown by Campbell and Valone plotted 0(0) for the capillary arrays, but F (0) for the efiirsive source. The result is a serious underestimate of the flux gradient for the effusive source. When the flux per unit area is considered, an effusive emitter placed 1/2 the sample diameter away has a flux at the edge of the sample that is only 25% of the flux at the center, not 71% as Reference 1 suggests. Such nonuniformity is unacceptable for most surface science experiments, particularly those such as infrared reflectance and temperature-prograrnmed desorption that average over the entire sample area. The enhancement factor E is given byI f,S 2n (1—sf,) kBT ’ 3'5 E21+ A where A is the sample area, S is the pumping speed, k3 is the Boltzmann constant, T is the temperature, s is the sticking coefficient and f; is the fraction of molecules leaving the closer that directly impinge on the sample. The right-hand side of Equation 3-5 gives the value of E if all molecules that do not immediately strike and stick to the surface are scattered into the vacuum chamber and subsequently pumped at speed S. E can be much larger if many of the molecules that miss the sample strike and are immediately trapped on other surfaces — cold surfaces of the sample mount or manipulator, for example. Since the expression for E involves quantities that are not properties of the doser, we 62 characterize the effective enhancement of the closer with fs, which is a property of the closer geometry alone. 3-2 Calculations The following computer modeling assumes a circular sample of diameter D parallel to a circular doser array of diameter d and a distance i away, as shown in Figure 3-1. The doser-to-sample distance is characterized by the angle 0,“, from the center of the closer to the edge of the sample; tan 0",“ = D/2l’. The capillary arrays consist of 5096 capillaries with L/a = 40, on a square net. Tests with larger numbers of capillaries showed that this was a sufficiently dense grid to be effectively continuous. The flux is calculated along a line of 50 points from the center to the edge of the sample. The effusive arrays consist of a set of holes equally spaced on a circle of diameter d, with or without an additional hole in the center. More complex geometries, with multiple rings of holes, would probably improve the results when the doser and sample are very close together, but do not provide much improvement for emax < 60°. Since the effusive arrays we considered have four-fold, but not fill] rotational symmetry, we calculate the flux at 1254 points in a square net covering one quadrant of the sample. At each point on the sample, the flux per unit area from the entire array, 00,,(5), is calculated by summing the contributions from all the individual sources: Garr(5) = 26(91') 3-6 63 where 0, is the angle from the ith source to the point I) and the sum runs over all the sources. Each .emitted molecule is given only one chance to stick so that no multiple collisions are included. We characterize the uniformity of the flux distribution with the area-weighted fractional standard deviation in flux, which is given by 9 3-7 linens-Gale 0’ = Gave where 1 .. Gave = '2 ij(P)dA 3-8 is the average flux, A is the sample area, and the integrals are taken across the entire sample. Since the flux distribution functions are normalized to the .total rate at which molecules are emitted, the fraction of gas directly hitting the sample is given by 3-9 64 where N is the total number of sources. All calculations are accurate within :l:1%; convergence was checked by performing calculations with larger numbers of data points. 3-3 Results and discussion Table 3-1 shows the values of f, and o for several effusive array designs with d/D = 0.8 and em, = 40° and 60°. Table 3-1. Comparison of circular effusive doser arrays with various numbers of holes. All designs except those with eight holes comprise a single center hole surrounded by a ring of equally spaced holes at d/D = 0.8. The eight-hole arrays have only the ring of holes, without a center hole. It can be seen that increasing the number of holes from 9 to 17 has little effect. The center hole improves the enhancement factor Slightly; it improves the uniformity for em, = 60° and degrades it for 0",“ = 40°, but in both cases the effects are small. 9m, Number of holes f3 0' 40° 5 0.32 0.15 9 0.31 0.12 17 0.30 0.1 l 8 0.30 0.09 60° 5 0.56 0.25 9 0.54 0.10 17 0.53 0.08 8 0.51 0.13 65 The five, nine and l7-hole arrays include a center hole; the eight-hole arrays lack the center hole. For 0”,“ = 60°, the inclusion of the center hole improves both the uniformity and the enhancement modestly; for 0",” = 40° it slightly improves the enhancement at some cost in uniformity. Increasing the number of outer holes from four (five-hole array) to eight (nine-hole array) leads to significant improvement, but further doubling the number of outer holes to 16 (17-hole array) has little effect. Based on these observations, we confine our study of effusive arrays to nine-hole arrays comprising a ring of eight holes around a central hole. Figure 3-2 compares the flux uniformity across the sample for several choices of doser design at different doser-to-sarnple distances, parameterized by the angle 0",“, (see Figure 3-1 for a definition of 0",“). The plots Show the flux per unit sample area Gm(p), normalized to the value in the center of the sample, 00,,(0), along a radial line. For the effusive arrays the line passes directly in front of one of the perimeter holes. The data are plotted versus the square of the distance from the center of the sample, p2, so that equal intervals along the x-axis correspond to equal areas of the sample. This presentation provides a more accurate visual representation of the uniformity than a plot of the flux VS. p, which underemphasizes the significance of the region near the edge. For those capillary array geometries that were also considered in Reference 1, the two calculations are in excellent agreement. For the single effusive doser, however, there is strong disagreement, since Reference 1 plotted the flux per solid angle F (0) rather than the flux per sample area G(0). Our results Show that the single effusive doser provides reasonable uniformity only when the doser—to-sample distance is very large (I > D). 66 1.0 %- l ‘5‘?--‘-""--, 0.9 ~ \‘35‘\ i * - \ 5 \\\ t , h _, \:,.\c \ag‘\-: 0.8 ~ V; 0.8 - . \ 09 f“? -\ ‘ . I“. 3: 0.7 *- ' \3\ E 0.7 s ' ‘ ‘ A ' .- s’ ’5 0.8 r . 3 0.8 - r ‘6 ' i U i s: > 0.5 ~ . > 0.5 r '3 a i i 3 04 i c 0-4: 0m,=25° ‘ u - . - 0.3 -T - 0.3 ~ . - -- - d/D = 2.0 capillary . . 0.2 I --0-- d/D = 1.0 capillary 1 0.2 . ,i ~2— dfb =1.0 effusive , . 1 ,2 —-?- d D = 0.5 effusive , 0 1 , . 0‘1 i —9— single effusive , ' L 9m“ '-" 40° , . A 0.0 ’ t 0”0.0 0.2 0.4 0.8 0.0 1.0 0.0 0.2 0.4 0.8 0.8 1.0 92 92 1.0 Wasn‘t-”run.” we. ‘ ’ 1‘ “ ‘ O. t ‘ 0.9 m\ V \ . ‘ .\ //‘\-11 film 1 ~ ‘9” . 0.8 r) g V. . - s \ /O’ \ 1 \ \e—«x/ . 0.7 r- , \\ '. r t \ X ‘ t 8 0-5 ’ E] \ ‘ 8 v , \ \ . , v u \\ . o > 0.5 " \\ "\~ 7 > m * ’. R * 3 F; 0.4 1- \ ‘ u 1 (Si . < 0.3 " \ V '1 L "is 5‘ 0.2 .\ \ * Bu °" i err... = 60° \r» 4. 0.0 . m . . J 0.0 0.2 0.4 0.6 0.8 1.0 p2 Figure 3-2: Normalized flux per unit sample area versus the squared distance from the sample's center for various sample to closer distances, indicated by 9max = arctan(D/2€). d/D is the ratio of the doser's radius to the sample's radius. Shown are plots for a capillary array of d/D = 2.0 and 1.0, a centered circular nine hole effusive array of d/D = 1.0 and 0.5, and a single effusive doser. 67 Not surprisingly, the greatest uniformity is provided by a large diameter (d/D = 2) capillary array. Perhaps more surprising is the relatively poor uniformity given by a capillary array equal in size to the sample. Even for 0mm = 75°, with the closer very close to the sample (3 = 0.13D), 20% of the sample area receives a flux that is smaller than that at the center by more than 20%. When the doser is farther away, 0mm < 45°, the nine- hole effusive array with d/D = 1 provides a more uniform distribution than the d/D = 1 capillary array. As the closer and sample are moved closer together the discrete nature of the effusive array becomes apparent in the strongly peaked flux distribution. It is likely that greater uniformity for short doser-to-sample distances could be achieved with more elaborate patterns of effusive emitters, but this is really the regime in which capillary arrays afford the greatest advantage. A complete comparison of closer designs must of course consider enhancement as well as uniformity. Figure 3-3 plots 0, the area-weighted fractional standard deviation of the flux, vs. f,, the fraction of emitted gas that impinges directly on the sample, for effusive and capillary arrays of varying diameters and for different values of 0”,“. (Our values of fs for the d/D = 1 capillary array agree with those of Campbell and Valonel and Winkler and Yatesz) The single effusive doser represents the d/D = 0 limit of the effusive array. In the d/D —-> 00 limit the value of a must again approach the single source value, since the outer holes are too far away to contribute any flux; this is why the effusive array curves have a "U" shape. 68 0.7 9 --O-- capillary array ’ . 0.6 ~ —t— 9 hole effusive array -- - I single effusive doser I - A smart = 400 0.5 — , .. . ., a 0.4 . —~ ~ D ° em“ = 25° . . . . 0.3 ~ -. o A t . ‘ 0.2 r "’ , ‘1 ; f .9 d/D = 1.0 Od/D=I.0 _d/D=260 “ . _ 0.1 - f . ‘ . . A g . x .. ’A’. ‘4‘. 0'000 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8 1.0 f, fs . . I 2.0 : ‘1' 9| ‘ i o h i e - 76° 3 1.6 L email = 60 4- \1 max - J . .. l“. . t : 1 1.2 " -‘ \i\ g. .1 D ' 41 A\\ l; i .' . ‘ 1’.‘ . ' a 1 ~\ . 0'8 F ' +5 V ‘ d/D 0 ' d p o .‘ g l. l‘- t a d/D = 2.0 f P .. k..\ > I} o \ “‘7 . . 0.4 i' \x f o a. r'" _, r \ . -‘ t. “ ‘ .. d D = 1.0 ' t\t.”‘,‘d/D=r.o J / ’--O 00i -A..A.-." " 1 AA AA-""—." . ' 0.0 0.2 0.4 0.8 0.8 0.0 0.2 0.4 0.6 0.8 1.0 f, fs Figure 3-3: Area-weighted fractional standard deviation 0 versus the fraction of emitted molecules directly impinging on the sample f, for various sample to closer distances, indicated by 9max = arctan(D/2l’). d/D is the ratio of the doser's radius to the sample's radius. The data shown are for a capillary array of d/D = 0.4, 0.6, 0.8, 1.0, 1.2, 1.4, 1.6, 1.8, 2.0; a centered circular nine hole effusive array of d/D = 0.3, 0.5, 0.65, 0.8, 1.0, 1.35, 1.7, 2.0; and a single effusive emitter. 69 An ideal doser would have a small value of o and a large value off,, and would lie in the lower right corner of the graph. Acceptable maximum values of o and minimum values off, of course depend upon the application. Several conclusions can be drawn from Figure 3-3: 1. Single effusive dosers and small capillary arrays (d/D << 1) are of little use unless very large flux variations (0 > 0.2) are acceptable. 2. For any value of 0mm capillary arrays outperform effusive arrays. 3. For large doser-to-sample distances (small 0",”) dosers of any design are of marginal usefulness, since it is virtually impossible to achieve f, greater than 0.3. 4. For small doser-to-sample distances, capillary dosers with diameters 20 - 40% larger than the sample perform extremely well. 5. Effusive arrays perform best when the holes are placed near the perimeter of the sample. 6. For intermediate doser-to-sample distances (0m, ~ 60°), effusive arrays perform nearly as well as the best capillary arrays. Under some experimental conditions these conclusions must be modified. One reason for using a doser arises when the dosed molecules are very rapidly pumped, for example by cold surfaces in the chamber, so that backfilling requires admitting an excessive amount of gas. In such cases a large enhancement E can be achieved even with a small value of 1;, and an effusive doser placed far enough from the sample to ensure good uniformity becomes a more attractive option. 70 A second situation in which dosers are often used is when the dosed molecules have a low sticking probability s. In this case the flux uniformity for capillary arrays placed close to the sample will not be as good as our calculations suggest. Molecules that do not stick to the sample are likely to scatter off the closer or its mount and reimpinge on the sample. These multiple collisions, which are not included in our model, will tend to enhance the relative flux at the center of the sample at the expense of the edges, leading to a less uniform flux distribution. In such a case a larger doser-to-sample distance may be preferable, and an efiirsive array, because of its simplicity and smaller optimal diameter, becomes a more attractive option. One virtue of effusive arrays is that they are easily made, and can be tailored to unusual shapes and sizes of samples. For example, surface infrared Spectroscopy usually requires relatively large oval or rectangular samples, and the doser must often be placed at some distance from the sample to allow optical access. Since the infrared beam usually probes the entire sample surface, good uniformity is essential. An effusive array with holes placed near the edges of the sample can be an effective doser for such an application. The model presented here can be readily implemented using commercial software on a personal computer, and calculations can be performed for any desired sample and closer geometry. It is then straightforward to optimize a closer design for the demands of any particular experimental situation. 71 References lC.T. Campbell and S.M. Valone, J. Vac. Sci. Technol. A 3 (1985) 408. 2A. Winkler and J.T. Yates, Jr., J. Vac. Sci. Technol. A6 (1988) 2929. 3HOW. Beijerinck and NR Verster, J. Appl. Phys. 46 (1975) 2083. 4D13. Kuhl and R.G. Tobin, Rev. Sci. Instr. 66 (1995) 3016. 5). Benziger and RJ. Madix, Surf. Sci. 94 (1980) 119. Chapter 4 Infrared Reflectance Change Induced by CO Adsorption on Pt(111) Measurements of the broadband IR reflectance change AR/ R were made for CO adsorbed on a Pt(111) surface. The frequency, temperature, and CO coverage were systematically varied in order to test the predictions of the Persson-Volokitin scattering "2’3 The choice of the CO/Pt(111) system was made with several important model. considerations in mind. First, CO adsorption on Pt(1 1 1) has been extensively studied so that phenomena such as site occupationf’s”6 coverage dependence of the C=O stretch vibration,7 and ordering in the adlayer,8'9 among others, are well known. Second, the scattering model is developed assuming the metal’s electrical conductivity is well- described by the Drude model. Since Pt is a transition metal, it is not clear that the Drude model should apply in this case. Thus, testing the scattering model’s predictions for Pt should provide an interesting test of the extent of the model’s validity. Third, temperature dependence enters the scattering model through the ratio of the mean free path to the skin depth, [/8 , which will be discussed below. Pt varies over an interesting range of [/8 at accessible temperatures. The predictions of the scattering model rely on a knowledge of the conduction electron density n of the substrate. Table 4-1 lists a number of material parameters for Pt calculated from n, with n determined using two different methods. One set, labeled 72 73 “DC,” is obtained by assuming it corresponding to one electron/atom, and a value of the bulk scattering time t B inferred from literature values of the DC resistivity.lo The other set, labeled “optical,” is obtained from the Drude parameters that best fit optical properties measured at room temperature.ll The value of r B at 90 K for the optical set is estimated by assuming the ratio between 13(273 K) and 13(90 K) is the same as for the DC parameters.lo Our major results do not depend strongly on which set of parameters is used. Table 4-1. Bulk material parameters for Pt. Frequencies in s'1 have been divided by 27m to convert to cm". 03p to. 8 13' (273 K) [(273 K) [(90 K) 1cm") cm") (A) cm") (A) (A) DC3 77,000 372 207 968 79 330 Optical” 41,500 133 380 558 91 380 aElectron density based on one electron/atom; scattering time based on DC resistivity.lo bRoom temperature values from optical data,ll 90 K values calculated by assuming the ratio between 13 (273 K) and 1:,(90 K) is the same as at DC.lo The results discussed in this chapter were previously reported in a paper by Kuhl, Lin, Chung, Wang, Luo, and Tobin.’2 74 4-1. Frequency Dependence Frequency-dependence in the Persson-Volokitin scattering model takes the form of safe 4), 4-, R a) where a) is the frequency of interest, [ is the elastic mean free path, and 8 = c/a) p is the ' classical skin depth. 0) , is a characteristic rolloff frequency given by 0),: p, 4-2 where vF is the Fermi velocity, 0) pis the plasma frequency, and c is the speed of light. Since [, 8 , and (al are each determined by n, a property of the substrate independent of the adsorbate, the function j: which contains the entire frequency dependence, is a universal function for a given substrate. The effect of the adsorbate enters through the prefactor a which is given by a = m , 4-3 4c cos0 75 where 0 is the angle of the incident IR radiation and p is the Fuchs specularity parameter (i.e. the probability of specular scattering). In Equation 4-1 the universal function f is expected to approach one at high frequencies, so, while f determines the shape, a determines the magnitude of the absorption. As shown in Chapter 1, the complete analytical expression for AR/ R is ( I 95_ .40 (tn/01)!)2 Re[g(a)/a)l,[/5)] 44 R - 1:2 4a) dq ’ (l—[nmpcose ImJ:>8(a)/0),,[/8,q)) where g(a)/0) l,[/8) and 8(0) /a),,[/8,q) are complex-valued functions. Near and below 001 nonlocal effects become important. Equation 4-4 has been successfully applied to a number of experimental studies, including CO on Cu(100),1 O on Cu(100),13 and CO on Cu(111).'4 Figure 4-1 shows my measurements of AR/ R vs. frequency, for 0 = 0.33 ML CO on Pt(1 1 1) at room temperature. The coverage 0 = 0.33 ML was chosen because the maximum AR/ R occurs near there (see Section 4-3) and thus AR/ R is insensitive to coverage in that region. Additionally, larger changes are easier to measure. The data points below 700 cm'1 represent averages of several measurements at nearby frequencies. It is apparent that the data exhibit scatter larger than the (rather conservative) experimental error bars. This scatter does not arise from instrumental error in individual 76 measurements, but rather from run-to-run variations in the actual reflectance change, possibly due to impurity levels below the Auger sensitivity level. AS a result of this scatter, and the small magnitude of the reflectance change, only a very rough comparison with theory is possible. 0.000 I ' I ‘ I 0.33 ML CO/Pt(l 1 l) 315 K I -0.001 -0.002 l a l r I A l 0 500 1 000 1 500 2000 25 00 3000 -0.003 Frequency (cm‘l) Figure 4-1. Fractional reflectance change AR/ R as a function of frequency, for 0.33 ML CO on Pt(1 11) at room temperature. The lines represent fits of the Persson-Volokitin model1 to the data, with the specularity parameter p as the only adjustable parameter. The dotted line represents the best fit using DC parameters, and the solid line the best fit using optical parameters (see Table 4-1). 77 As seen in Chapter 1, the function g(a)/a) , , [/8) in Equation 4-4, given by g = idyil - ijF’(y)+ a): 4mm [idyi-l— - L] FUJI, 4'5 . y y’ -w’-iwn y2 y4 contains information about an expected antiabsorption resonance feature associated with a frustrated vibrational mode of CO on the Pt(111) surface at a frequency a) o and with a damping coefficient 11 . Our measurements did not probe a sufficiently low frequency regime to observe this feature. Furthermore, the elements of the Persson-Volokitin scattering model which relate the broadband decrease in AR/ R to the damping of . . . 15-17 hlndered modes remain controversral. For calculations of just the broadband reflectance change AR/ R, it is sufficient to use only the first term in Equation 4-5, thereby dropping all information about the antiabsorption resonance. Note that doing this ignores the claim by Persson that 1] cc (1 - p), which is embedded in a. Nevertheless, this approach has been successful fitting the experimental AR/ R data in the study by Lin et al. ’3 of O on Cu(100). Under this approach, once a)l is calculated from the conduction electron density, Equation 4-4 contains only one adjustable parameter: the magnitude of the asymptotic reflectance change a. The dotted line in Figure 4-1 shows the best fit to the Persson-Volokitin model achievable using the DC material parameters to determine a)l . The solid line shows the best fit using the optical parameters. The data are consistent with the model, especially 78 when the optical values of the bulk parameters are used. Unfortunately, the uncertainties both in the data and in the material parameters preclude any firm conclusions. 4-2. Temperature Dependence Within the Persson-Volokitin scattering theory, the only parameter that is expected to show a strong dependence on temperature is the mean-free-path [ = th B , which enters through the ratio [/ 8 in the luliversal function f. The scattering time ‘t 8 increases strongly with decreasing temperature, while VP and 8 are' essentially temperature-independent. As can be seen from the DC values given in Table 4-1, for Pt [ changes by more than a factor of four going from 79 A at 273 K to 330 A at 90 K (our measurement temperature, close to liquid nitrogen temperature). Since no optical measurements leading to Drude parameters are available at temperatures comparable to that of liquid nitrogen, we assumed the optical parameters scale like the DC parameters. This gives a corresponding change in [ from 91 A at 273 K to 380 A at 90 K Similar to Figure 5 of Reference 1, Figure 4-2 shows a number of curves calculated from the Persson-Volokitin scattering model. The three quantities shown are plotted against the ratio [/8, which is a scale related to the inverse of the temperature. Each curve represents a different frequency regime, parametrized by the ratio 00/00l . The first set of curves, Figure 4-2(a), illustrates the absolute change in reflectance AR. Note that the vertical scale in Figure 4-2(a) is a negative scale, with zero at the top, because AR is an absorption of the IR light. For [ / 8 greater than about two (i.e. toward low temperatures), AR depends only very weakly on [/8, but for lower values (i.e. toward 79 higher temperatures) AR increases rapidly with increasing [ / 8 (decreasing temperature). For Pt we estimate [ / 8 = 0.4 (0.24) at room temperature and 1.6 (1.0) at 90 K using DC (optical) parameter values. We would therefore expect |AR| to increase by about a factor of two when the sample is cooled from room temperature to 90 K, depending slightly on the measurement frequency. The experimentally measured quantity, however, is the fiactr’onal reflectance change AR/ R, where R is the clean-surface reflectance, which also increases’ with increasing [/ 8, as shown in Figure 4-2(b) (the details of this curve depend on 11; Figure 4- 2(b) is calculated using the DC value for Pt). This effect tends to offset or even overcome the change in AR. In fact, as shown in Figure 4-2(c), the magnitude of the fractional reflectance change AR/ R actually decreases with increasing [/ 8 for 00/0) 1 > 4.6. It is worth noting that the importance of the temperature dependence of R to the ratio AR/ R has not been previously recognized. In fact, the failure to account for this dependence actually led Persson to an incorrect conclusion regarding the AR/ R for CO on Ni(100).l 80 Figure 4-2. Dependence of the reflectance on [/8 predicted for adsorption on Pt by the Persson-Volokitin scattering theory,l for the indicated values of ctr/0)l . The vertical scale for (a) and (c) is arbitrary. The small arrows indicate the room temperature (RT) and 90 K values of [ / 8 for Pt. DC values of the bulk parameters have been used (see Table 4-1). (a) Absolute reflectance change AR. (b) Clean-surface reflectance R. (c) Fractional reflectance change AR/ R. 81 . _ _ q a a q q a a _ _ _ a A a q a 4 q a q q a a T a l l t J \I .a 44 ) \l 4 a a.“ L2 (av . f O E . .a 4? e o t _ T L I o— A .. _. a *0. c a #0 _ c i. I 1 TI — lo. M ... . e K x p. 2 9 . i a w A. m 0 .0 O — ~ . » r __ l .. a4 + A. b a 40 .I. liii l l.Iu—I|‘Ii4 i ii i i i g iii .fi 6 O. l l 1 A P. 15 . .. .. .l J r a .a a 0 ~. I m 4. A . .. . a 6.0 O O L . r r ,./ i r .. lo. .. .. I. a a 0.9 if a 40 T e. , o E, l I R C D T . . l T ./‘ R I. R . l r 1 r 5 + a I O + I . Fir/ll 1+ + a I 9 o . 1 .II 1 i illl liiiii lbril. 4’1, IL 0 W4 / a a a O .a rut. k a a ,0 0 0 0. 0. o. 0.0.0. - 0 0. 0 o. L 4 7 N I 47m L 4 7 m > f p E p p _ p L E — p p - r p P P . h b b r W p P p o. 0 2. 4. 6. 8. 0 0. 9 8. 7. 6. 50 4. s a 6. 0.0 0 o D a mu 1.. I 0 0 0 o 00 o o 4. .... 9.. 2.3 so e :5 ~52 82 In order to perform the first experimental study of the effect of changing [/ 8 for a single adsorption system, I made a number of measurements at 90 K at frequencies of 2500 cm'1 and 2800 cm", corresponding to c0/0), values close to 7 (20) for DC (optical) parameters. Over a range of CO coverages near 0.33 ML, the largest value of AR/R observed was -0.20%i0.02% compared with -0.28%d:0.03% at room temperature, giving a ratio of the 90 K to room temperature AR/ R of 0.73:0]. The theoretically predicted ratio using DC (optical) parameter values is 0.85 (0.72). In View of the uncertainties in both the calculation and the experiment, we regard this as reasonable agreement. It would be of interest to perform similar measurements at 0) /0)l < 4.6 , but we have not yet attempted them. My measurements also permit an estimate of the ratio of the clean surface reflectance at 90 K to that at room temperature; we find R(90 K)/R(315 K) = 1.3i0.02. The predicted ratio, using DC (optical) parameter values, is 1.2 (1.5), so the agreement between experiment and theory is again good. 4-3. Coverage Dependence For a dilute, disordered overlayer, the scatterers will be independent and the scattering of electrons will be incoherent — the scattering rates from different adsorbates will simply add. The scattering probability can then be described by a scattering cross section per adsorbate 2, (l—p)= n02, 4-6 83 where It, is the surface density of adsorbates — the coverage. If the scattering cross section per adsorbate is constant with respect to coverage, the resulting reflectance change (or resistivity change) will vary linearly with adsorbate coverage. This can be seen by substituting Equation 4-6 into the expression for the asymptotic value of AR/ R, Equation 4-3, which gives a- 3vF-2 — 4-7 4c cos0 a O A linear dependence is in fact generally observed at low coveragem’m'25 For CO on Cu(100) and Cu(111) the linear dependence persists up to saturation.22’23 For other systems, however, nonmonotonic dependences have been observed. The first measurements of nonmonotonic coverage-dependent reflectance changes were made by Riffe et al.2"26 They measured the change in attenuation of surface electromagnetic waves (SEW), which is the inverse of measuring the change in IR reflectance, but gives similar information. They studied N2, 02, CO, H2, and D2 on W(100). The magnitude of the SEW attenuation coefficient 01 increased monotonically with coverage for N2, while for both 02 and CO it reached a maximum and then dropped at higher coverages. H2 and D2 showed a more complicated nonmonotonic dependence. In a subsequent IR reflectance study of H on W(100), Riffe and Sievers correlated the nonmonotonic coverage-dependence of the reflectance change with ordering of the 84 overlayer, at a frequency where the reflectance change was dominated by scattering rather than electronic state effects.” This correlation is based on adlayer patterns observed in separate low energy electron diffraction (LEED) experiments.28 Another example of a nonmonotonic coverage dependence in AR/ R was reported by Lin, Tobin and Dumas for O on Cu(100).13 The magnitude of the reflectance change increased monotonically up to 0.25 ML, but decreased by about 20% when the coverage increased to 0.35 ML. While only four coverages were studied, the departure from the low-coverage linear behavior at the highest coverage was well outside the experimental uncertainty. Similar to Riffe et al., Lin and coworkers tentatively attributed their observations of a nonmonotonic coverage-dependence to ordering of the overlayer. It is well known from the literature that 0 forms a (2J2 x J2)R45° structure with a “missing row” reconstruction (every fourth row is vacant) on Cu(100) at high coverages.”34 Using LEED, Wuttig, Franchy, and Ibach established29 that there is a first order phase transition, at and above room temperature, from a low-coverage disordered phase into the ordered phase at a critical coverage of BC = 0.34. Note that this is very close to the coverage at which Lin et al. saw a drop in AR/R. Unfortunately, when they attempted LEED observations themselves they were unable to correlate ordered LEED patterns with the IR measurements due to temperature constraints. The idea that ordering of the overlayer could be responsible for a nonmonotonic coverage-dependence is based on a reduction of the extent of nonspecular scattering. A fully ordered overlayer restores translational symmetry to the surface (the scattering from different adsorbates is coherent) and diffuse scattering is no longer possible. Nonspecular 85 processes are allowed only if the scattering vector is a reciprocal lattice vector of the overlayer. Since these vectors are usually different from those of the clean surface, and since the scattering amplitudes are in any case different, the reflectance need not return to its clean surface value.l4 Nevertheless, it can be expected to be different from the reflectance from a disordered overlayer of the same surface density. Hirschmugl and coworkers, on the other hand, claimM’ZZ’23 that ordering in the CO on Cu(100) and Cu(l 1 1) systems has no effect on the coverage dependence of AR/ R. There are ordering transitions in both of those systems and a linear coverage dependence was nevertheless observed. Thus it was with the following questions in mind that we investigated the coverage dependence of CO on Pt(111): ( 1) Is the coverage-dependence for CO on Pt(1 11) monotonic or nonmonotonic? (2) If nonmonotonic, does ordering in the overlayer explain the coverage-dependence? (3) If not ordering, what mechanism accounts for a nonmonotonic dependence on coverage? Figure 4-3(a) shows the fractional reflectance change AR/ R as a function of CO coverage for CO on Pt(111) at ~ 315 K at both 2500 cm'I and 2800 cm". These frequencies were chosen to be high enough that AR/ R is frequency-independent, and well away from the C=O stretch vibrations at 1850 and 2100 cm". Within the Persson- Volokitin scattering model AR/ R becomes frequency-independent for 0) >> I g‘ ,0) I. As shown in Table 4-1, 1 3'1 corresponds to at most about 970 cm", while 0)l is considerably lower. 86 0.0000 . . . . . . . r - r .- ‘ j ‘1 -00005 - _f E _._ 2800 cm" : - ___=—__ ; —:-—- 2500 cm" i r .L L ‘ . i I ' —0.0010 _ t——:—.-——+ " l. I — i' ' ° - r —s— r % -00015 - _a—: T ‘ Dfi ‘” —!—— . T . Q ' =- T 4 ‘ l I 00020 - )———I_.' ' __._t__.=‘:_:_1 .. i. .. +r—-—-1: ' -1 i—O—_-___a__j _'__ l ' I“ ' - - r—z—l ~ ' - 0.0025 . i—T . ' szT— . 1 00030 - (a) I r 1 1 L I r l A r l r I r F fl 1 I r 0.00 r- ‘ - A 004 - ' - o 1- ‘ a A a T) -0.08 - ‘ a :5 <1 ' e‘ a ‘ -0.12 . o ‘ T a A -0.16 - .A " . (b) . -0.20 1 1 ' ' ' ' ' ' ' ' 0.0 0.1 0.2 0.3 0.4 0.5 CO coverage (ML) Figure 4-3. (a) Fractional reflectance change AR/ R as a function of coverage, for CO on Pt(1 l l) at 315 K. Data are shown for two 1R frequencies, 2500 and 2800 cm”. (b) Work function change for CO on Pt(1 l 1), reproduced from References 35 (Norton, et al., triangles) and 36 (Ertl, et al., circles). 87 Strong nonmonotonic coverage-dependence was Observed for CO on Pt(1 11). The magnitude of AR/ R increases with increasing CO coverage up to about 0 = 0.33 ML. Above 0.33 ML the magnitude of AR/ R decreases with increasing coverage until saturation is achieved at 0.5 ML. It is interesting to note that 0.33 ML is the coverage at which the CO overlayer forms a (J3 x J3)R30° ordered structure on Pt(1 l 1), as observed9 by LEED. Below 0.33 ML CO forms a disordered overlayer on Pt(1 11). At 0.5 ML the overlayer forms a c(4 x 2) ordered structure.9 It is tempting, therefore, to attribute the decrease in [AR/RI to ordering, following the suggestions made for H on W(100)27 and O on Cu(100).’3 Since long-range order is not achieved at room temperature, however, ordered LEED patterns cannot be observed for CO above 300 K. It was necessary to cool the sample in order to achieve observable LEED patterns. In order to cool the sample, it was necessary to fill the coldfinger with liquid nitrogen. During a complete IR measurement a sufficient quantity of liquid nitrogen boiled away for the liquid level in the coldfinger to drop significantly. This allowed some fraction of the coldfinger’s length to begin to warm up and lengthen. The lengthening of the coldfmger caused minute changes in the sample’s position, which rendered the IR measurement unusable. Thus, systematic coverage-dependent measurements were not possible below room temperature, and we have no direct evidence for adsorbate ordering. If ordering in the overlayer is the reason for the nonmonotonic coverage dependence at room temperature, the effect must be assigned to partial, short-range ordering, for which there is no direct experimental evidence. 88 An additional explanation for a nonmonotonic coverage-dependence could be that the scattering cross section per adsorbate varies with coverage, either because different binding sites are being occupied or due to coverage-dependent changes in the electronic structure of the adsorbates. It is well known that CO occupies both atop and bridging sites on Pt(1 1 1) and that 445.35 Moreover, the two the relative populations of the two sites are coverage-dependent. species have distinctly different electronic states and dipole moments,35 so it is entirely plausible that they might have different scattering cross sections for conduction electrons. Site exchange alone, however, cannot account for the drop in |AR/R| seen in Fig. 4-3(a). The largest possible drop from this effect would occur if the layer were entirely atop- bonded at 0.33 ML and half atop, half bridging at 0.5 ML, and if the bridging species had zero cross section. In this case we would have: 1%(050 ML) _ 0.25 _ _ —0.76, 4-8 %(0.33 ML) 0.33 ( ) whereas the experimental value of the ratio is 0.43. Moreover, it is clear that bridge Sites are occupied even at 0.33 ML and that the atop population drops very little, if at all, between 0.33 and 0.50 ML.5'35 The final possibility that we consider is a coverage-dependent scattering cross section caused by changes in the chemical bonding of the CO. Strong evidence for such changes comes from measurementsms’36 of the work function change All), which show a 89 coverage dependence strikingly similar to that of AR/ R, as shown in Fig. 4-3(b) where the work function data of References 35 and 36 are reproduced. (It should be noted, however, that the coverage scale for Reference 35 was set by assuming that the minimum value of Ad) occurs at 0.33 ML; in Reference 36 the coverages were determined directly from temperature programmed desorption (TPD) and LEED.) The rapid increase in All) between 0.33 ML and 0.4 ML or 0.5 ML is attributed"'35‘36 to depolarization of the adsorbate bond, which could very well also be accompanied by a decrease in the scattering cross section. We currently regard this as the most likely explanation for Figure 4-3(a), but the ordering hypothesis cannot be ruled out. The data in Fig. 4-3(a) can also be used to estimate a value for the cross section 2 for diffuse scattering of Pt conduction electrons from the CO adsorbates. Equation 4-6 gives a linear expression relating the asymptotic value of AR/ R to the coverage in the form of the surface density of adsorbates n“. We measured a linear relationship for coverages below 0.33 ML. The scattering cross section 2 can be extracted from the slope of that line. Using DC (optical) values for the material parameters, we find 2 = 0.8 A2 (0.9 A2), for coverages below 0.33 ML. For comparison, Table 4-2 lists cross sections for diffuse scattering of conduction electrons for a number of other adsorbate systems, determined via measurements of the coverage-dependent adsorbate-induced change in resistivity. Compared to other chernisorption systems, CO on Pt(111) is apparently a rather weak scatterer. Also for comparison, the area per Pt atom on the surface is 6.6 A2 and the area per CO molecule at 0.5 ML coverage is 13.2 A2. 90 Table 4-2. Scattering cross section per adsorbate for several adsorption systems. system 2 (A2) reference chemisorption CO/Pt(1 l 1) 0.8 H/Ni(l l l) 8.2 3 CO/Ni(lll) 16.0 3 Nz/Ni(l 11) 4.8 3 CO/Cu(111) 5.8 3 O/Cu(111) 18.8 3 Ag/Ag(l 1 1) 14 37 physisorption HZO/Ag(1 l l) 0.06 2 CO/Ag(1 l l) 1.0 3 C2H4/Ag(l l l) 0.5 3 Xe/Ag(111) ~ 0.6 3 C6H6/Ag(l l l) 0.6 3 Can/Agfl 1 1) 0.6 3 CzHg/Ag(l 1 1) 0.1 3 91 44. Conclusions A number of conclusions can be drawn from these measurements of the adsorbate-induced change in the broadband IR reflectance AR/ R of Pt(1 11) as a function of frequency, temperature, and CO coverage. The Persson-Volokitin scattering model does a good job treating the temperature- and coverage-dependence of CO on Pt(1 1 l), a transition metal surface. This is significant because the the scattering model assumes that the electrical conductivity is described by the Drude free electron model. The Drude model is not expected to apply to transition metals. It is important to consider the temperature dependence of the clean surface reflectance R in addition to the temperature dependence of the absolute change in reflectance AR when dealing with the fractional change AR/ R. Further work must be done to understand the presence of a peak in the coverage dependence in selected systems. For the CO/Pt(11 1) system the peak cannot be attributed to changes in adsorption site, but could be attributable either to partial ordering in the overlayer or to changes in CO'S electronic structure. Broadband IR reflectance measurements can provide an accurate measure of the scattering cross section of conduction electrons from adsorbates and can offer a new probe of surface dynamics. Calculating and predicting the values and variations of these cross sections presents a challenging and fertile field for theoretical investigation. 92 References 'B.N.J. Persson and AI. Volokitin, Surf. Sci. 310 (1994) 314 . 2B.N.J. Persson, Chem. Phys. Lett. 197 (1992) 7. 3B.N.J. Persson, Phys. Rev. B 44 (1991) 3277. 4GS. Blackman, M.-L. Xu, D.F. Ogletree, M.A. Van Have, and GA. Somorjai, Phys. Rev. Lett. 61 (1988) 2352. 5W.D. Mieher, L.J. Whitman, and W. HO, J. Chem. Phys. 91 (1988) 3228. 6IV. Nekrylova, C. French, A.N. Artsyukhovich, V.A. Ukraintsev, and 1. Harrison, Surf. Sci. Lett. 295 (1993) L987. 7J.S. Luo, R.G. Tobin and D.K. Lambert, Chem. Phys. Lett. 204 (1993) 445. 8K. Horn and J. Pritchard, J. Phys. (Paris), Colloq. C4, Suppl. No. 10, 38 (1977) 164. 9H. Steininger, S. Lehwald and H. Ibach, Surf. Sci. 123 (1982) 264. loJ. Bass and K.H. Fischer, in: Landolt-Bornstein Numerical Data and Functional Relationships in Science and Technology, vol. 15a, ed. K.H. Hellwege (Springer-Verlag, New York, 1982) 63. llM. A. Ordal R.J. Bell, R.W. Alexander, Jr., L.L. Long, and MR. Querry, Appl. Opt. 24, (1985) 4492. 12D. E. Kuhl, K. C. Lin, C. Chung, J. s. Luo, H. Wang, and R. G. Tobin, Chem. Phys. 205 (1996) I, invited paper in special issue on “Surface Reaction Dynamics.” l3K.C. Lin, R.G. Tobin and P. Dumas, Phys. Rev. B 49 (1994) 17 273 (1994); ibid. 50 (1994) 17760. ”or. Hirschmugl, G.P. Williams, B.N.J. Persson and A. I. Volokitin, Surf. Sci. 317 (1994) L1141. 93 15R.G. Tobin, Phys. Rev. B 48 (1993) 15 468 ”B.N.J. Persson, Phys. Rev. B 48 (1993) 15 471 ”TA. Germer, J .C. Stephenson, E.J. Heilweil and RR. Cavanagh, J. Chem. Phys. 101 (1994) 1704. l8P. Wissmann in: Surface Physics, ed. G. H8hler, Springer Tracts in Modern Physics Vol. 77 (Springer, New York, 1975). '9D. Dayal, H.-U. Finzel and P. Wissmann in: Thin Metal Films and Gas Chemisorption, ed. P. Wissmann (Elsevier, Amsterdam, 1987). 20D. Schumacher, Surface Scattering Experiments with Conduction Electrons, ed. G. H8hler, Springer Tracts in Modern Physics Vol 128 (Springer, New York, 1993). 2'D.M. Riffe, L.M. Hanssen and A.J. Sievers, Surf. Sci. 176 ( 1986) 679. 22C]. Hirschmugl, G.P. Williams, F .M. Hoffmann and Y. J. Chabal, Phys. Rev. Lett. 65 (1990) 480. 23C.J. Hirschmugl, Y.J. Chabal, F.M. Hoffmann and GP. Williams, J. Vac. Sci. Technol. A 12 (1994) 2229. 24E. Borguet, J. Dvorak and H.L. Dai, SPIE Proceedings 2125 (1994) 12. 25M Hein and D. Schumacher, J. Phys. D: Appl. Phys. 28 (1995) 1937. 2(’D.M. Riffe, L.M. Hanssen and A.J. Sievers, Phys. Rev. B 34 ( 1986) 692. 27D.M. Riffe and A.J. Sievers, surf. Sci. 210 (1989) L215. 28D.A. King and G. Thomas, Surf. Sci. 92 (1980) 201. 29M. Wuttig, R. Franchy, and H. Ibach, Surf. Sci. 213 (1989) 103. 3"BC. Zeng, R.A. McFarlane, and K.A.R. Mitchell, Surf. Sci. 208 (1989) L7. 31l.K. Robinson, E. Vlieg, and s. Ferrer, Phys. Rev. B 42 (1990) 6954. 32MC. Asensio, M.J.Ashwin, A.L.D. Kilcoyne, D.P.Woodruff, A.W. Robinson, Th. Lindner, J.S. Somers, D.E. Ricken, and AM. Bradshaw, Surf. Sci. 236 (1990)]. 94 33Ch.Woll, R.J. Wilson, S. Chiang, H.C. Zeng, and K.A.R. Mitchell, Phys. Rev B 42 (1990) 11926. 3"R. Mayer, C.-S. Zhang, and K.G. Lynn, Phys. Rev. B 33 (1986) 8899. 35PR. Norton, J.W. Goodale and EB. Selkirk, Surf. Sci. 83 (1979) 189. 36G. Ertl, M. Neumann and KM. Streit, Surf. Sci. 64 (1977) 393. ”B.N.J. Persson, D. Schumacher, A. Otto, Chem. Phys, Lett. 178 (1991) 204. Chapter 5 Changes in the DC Resistivity and the Infrared Reflectance Induced by Oxygen and Formate Adsorption on Cu(100) It has long been known that adsorbates increase the DC resistivity of thin metal films."8 Attempts have been made to relate resistivity changes to measurements of adsorbate-induced reflectance decreases in the visible light region?“ More recently, sensitive measurements of the broadband absorption of IR light caused by adsorption on a metal surface”’13 has renewed interest in the possibility of a relationship between these two physical effects. Using a model based on conduction electron scattering, Persson, building on earlier work, developed a linear relationship14 between the adsorbate-induced resistivity change Ap and the adsorbate-induced broadband IR reflectance change AR/ R. As shown in Chapter 1, following the example of Lin et al.,15 the expression relating Ap to AR/ R is given by AR_ 4ne2 ) _R—_ (mccos0 [tAp] 5-1 where n is the conduction electron density, e is the electronic charge, m is the effective electron mass, c is the speed of light, 0 is the incidence angle of the IR radiation, and t is the thickness of the thin film sample. The only quantity in Equation 5-1 that refers to a 95 96 thin film is tAp; the rest of the terms refer to bulk properties. This is not an inconsistency, however, since the Persson scattering model is a slab model in which the near surface region of a bulk sample can be mapped to a thin film. The quantity tAp has been shown to be independent of film thickness.‘5 In this chapter I present experiments that were performed to test the linear relationship predicted in Equation 5-1. 0 adsorption on a Cu(100) film will be discussed in Section 5-1. Formate adsorption on the same Cu(100) film will be discussed in Section 5-2. 5-1. 0 Adsorption on a Cu(100) Thin Film Simultaneous measurements of the resistivity change and reflectance change were performed while dosing in situ grown epitaxial Cu(100) films with 02 and, subsequently, formic acid in order to test the linear relationship between AR/ R and tAp . The formic acid dose will be discussed in the next section. The Cu(100) films were grown on etched Si(100) substrates as described in Chapter 2, Section 2-4, and generally conform to the characterization described therein. With the help of E.T. Krastev, a number of runs have been performed which have established a high degree of reproducibility. For the purposes of this discussion I will restrict myself to the single run shown in Figure 5-1. The Cu(100) film studied in the time scan shown in Figure 5-1 was grown at ~ 40°C, 2 A/s, and a background pressure during growth of ~ 4.8 x 10''0 Torr. Its dimensions were 1.000 cm x 1.880 cm x 507 A. The resistance, measured ex situ, was 3.26 (2, which leads to a resistivity of p = 8.79uQ-cm. The value for bulk Cu is 97 p = 1.70uQ-cm .16 The combined resistivity and reflectance measurement, described in Chapter 2, Section 2-5, was performed with the spectrometer set at 2800 cm". The initial UHV background pressure was 1.8 x 10'lo Torr. After about 200 s of the resistivity and reflectance time scans had elapsed, the sample was dosed with 50 L 02 by backfilling the UHV chamber (note that the closer was not employed) to 5.0 x 10'7 Torr for 100 3. At about the 500 3 point, the sample was exposed to 50 L of formic acid vapor, using the same dose of 5.0 x 10'7 Torr for 100 5. At the end of an 02 or formic acid dose at a dosing pressure of 5 x 10'7 Torr, the UHV chamber typically reaches the low 10'9 Torr region in under 10 seconds. After that, it requires on the order of another minute to achieve a pressure in the mid 10'‘0 Torr region. Figure 5-1 shows the “raw” data. During the time scans shown in Figure 5-1 the coldfinger was full of liquid nitrogen and the sample was held at 28°C via the tungsten filament heater. Just before the experiment began, the coldfmger was refilled. This resulted in a damped vibration of the coldfinger which in turn affected the sample’s position, which showed up as erratic drift in the early part of the baseline of the IR time scan. Consequently, the first 24 s of data have been eliminated from both the IR and resistance time scans. A remnant of the drifi can be seen in the data prior to about 150 s in Figure 5-1(a), after which the drift throughout the rest of the time scan is essentially linear. 98 0.02220 ' T ‘ T V I V r ' T r ' I - . ! = L (a) . 0.02215 ~ M . : . A, - L ‘ ‘,__.' dosing . ”l dosing ‘ 5; 1501.02 ' :soLrormic ‘5 0.02210 ' l ‘ - 3CId "‘ '2' : . 2 . t .5 0.02205 - I i s ; I 3 . 5 3; 0.02200 A . - “no 2 . Cu(100) film , I 02 9 2800 cm‘1 I 0. 1 5 l- 313 K A i 4 0.02190 4 . 4 . A 1 L 0.00297 A A A A A . A A A b i (b) 1 0.00294 - _ g l . r Cu(100)f11m , g * 313 x 4 0.00291 - ' + E * ‘ 3 ». i , . 3 . j t , 3 0.00288 » i . g . g . F = f ‘ > » g i __ l ‘ 0.00285 - ‘ 3 ‘ ~ t f i ’ dosing ? . dosing 0.00282 ~ 1 l 50 L o, I 50 L formic - ‘ ' acid P ' i 4 0.00279 4 . 11 . 1 . L . 1] . ll . l i o 100 200 300 400 500 600 700 800 time (s) Figure 5-1. Time scans of (a) the reflected IR intensity (i.e. voltage output from a lock- in amplifier) at 2800 cm'1 and (b) the voltage drop across the sample (i.e. . voltage output from a lock-in amplifier) with an applied current. Both 50 L doses involved background pressures of 5 x 10'7 Torr for 100 s. 99 0.002 A . . , . , . , - f , . C ’ . i (a) . 0.000 - - -- .................. .................. _ -0.002 . . ‘ dosing l g 0‘0“ l' ’3 SOLforrnic 1 § acid ‘ -0.006 i . : Cu(100) film ; l 1 2800 car! - . -0.008 r 313 K | . i i -0.010 A . - A i A 0.06 - . . .1 A .1 . r . ,, - ,1 r , . ' ' f l (b) . 0.05 - 5 . Cu(100) film i 313 K ' 0.04 A ' . g 0.03 - - 0.02 - . . 0'01 ' . dosing ‘ dosing J _ SOLO; SOLformic acid 0.00 - . l A ll . l L O l 00 200 300 400 500 600 700 800 time (8) Figure 5-2. Time scans of (a) the broadband reflectance change AR/ R at 2800 cm'l and (b) the fiactional change in resistivity Ap/ p. Each curve has a linear drift removed. Both 50 L doses involved background pressures of 5 x 10'7 Torr for 100 s. 100 It may be noted that the vertical lines representing the beginning and ending of the dosing do not precisely line up between Figures 5-l(a) and 5-l(b). This is due to the fact the two data sets were collected on separate computers; the time axes begin close to but not exactly at the same point. The raw IR intensity and voltage drop data of Figure 5-1 is represented in Figure 5-2 as the broadband reflectance change AR/ R and the fractional change in resistivity Ap / p respectively. A linear drift has been removed fi'om each of the curves shown in Figure 5-2. For the IR reflectance data the slope of the curve was determined from the data between 120 s and 202 s because this is where the baseline appears to have stabilized, as can be observed by the consistency between the slope before and after dosing in Figure 5-l(a). The dashed line at AR/ R = 0 in Figure 5-2(a) is provided as an aid to the eye. For the resistivity data, the slope was determined from the data between 26 s and 202 s. The resistivity data had an extremely stable baseline, so that removing the linear drifi had essentially a negligible effect. It can be immediately observed from Figure 5-2 that there is a strong correspondence between AR/ R and Ap / p . At the beginning of the dose, the broadband IR reflectance AR/ R drops by about 0.9% while the resistivity increases by about 5.5%. It is interesting to note that both Ap/ p and [AR/RI appear to decrease somewhat after achieving a maximum. The value of Ap / p just before the formic acid dose is about 5.2% while AR/ R is about -0.8%. This could be related to a nonmonotonic dependence on coverage for AR/ R that was reported by Lin et a1. 17 for O on Cu(100) (see Chapter 4, 101 Section 4-3 for a discussion of nonmonotonic dependences of AR/ R on coverage). It is conceivable that both AR/ R and Ap/p have a nonmonotonic dependence on coverage here, although the fact that the decrease appears to continue after the dose when the coverage is not changing could point to some other explanation. At room temperature 02 is known to dissociate on the Cu surface and leave an 17-25 overlayer of atomic oxygen. One might suspect that the decrease in Ap/p and [AR/RI following the dose results from an oversaturation coverage of 0 during the dose followed by O leaving the surface as the pressure in the chamber drops. This is not a likely explanation for two reasons. First, the decrease clearly begins before the dosing is completed. Second, it is known”’30 that O has a very small sticking coefficient on Cu which gets smaller at higher coverages. We have no independent measure of coverage at this time, but it is likely that the coverage is well below saturation, which is 0.5 ML.20 Moreover, the O that does stick does not desorb thermally at room temperature.“30 The absence of ordered patterns observable by LEED at room temperaturen’zo'25 indicates that the decrease probably cannot be attributed to ordering in the overlayer either. Additional possible explanations include relaxation in the overlayer or adsorbate- induced surface reconstruction. A straightforward means of testing the linearity predicted by Equation 5-1 is to plot AR/ R from Figure 5-2(a) versus tAp extracted from Figure 5-2(b). Note that the slight difference in the origin of the time axes becomes significant here, however, the sharp onset of change observed in both Figures 5-2(a) and 5-2(b) allows for a convenient reference point. Figure 5-3 is a plot of AR/ R versus tAp for this experiment. 102 0.001 ’ .\ -0 001 L .\\\ \. \. .\\ 0003 ~ \ \\ -0.005 r - \ g -0.007 A \ \ oft... o -0.007 ~-0.008» ._';t; 0.009 l "I... N}; -0.009 - ' ° -°'_~. ‘ 0.010 5 2.30 2.36 2.42 2.48 _0011 1 1..-2 grAihiL - : 1 -0.1 0.2 0.5 0.8 1.1 1.4 1.7 2.0 2.3 2.6 tAp (10*5 unocmz) Figure 5-3. AR/ R versus tAp for 02 adsorption on a Cu(100) film. The inset is an enlargement of the lower right hand comer. The solid line is a linear fit to the data, which are taken from about the 200 s to 500 5 region of Figures 5-2(a) and 5-2(b). Figure 5-3 clearly exhibits qualitatively the linear relationship between AR/ R and tAp predicted by Equation 5-1. The data plotted in Figure 5-3 come from the 206 s to 500 8 region of Figure 5-2(a) and the 208 s to 502 5 region of Figure 5-2(b). Note that the 02 dose ends at about 300 seconds, so that ~ 2/3 of the data plotted in Figure 5-3 comes from the interval in which the UHV chamber was pumping residual 02. The data appears 103 to move back up the curve a little in this region and then bunch up vertically near p = 2.3 x 10“ uQ-cm2 ; i.e. AR/ R drifts slightly more than tAp . Equation 5-1 also makes a quantitative prediction about the relationship between AR/ R and tAp . The slope of the line should be given by d(AR/ R) _ 4ne2 d (tAp) — -( me 0050] ° 52 The angle of incidence 0 of the IR radiation is 86° (see Chapter 2, Section 2-2). The conduction electron density n can be calculated by assuming one conduction electron per 26'” which gives a value for Cu of n = 8.5 x 1022 cm'3. Using these values, along atom, with literature values‘6 of the constants e, m, and c, a predicted value of the slope can be calculated. Equation 5-3 gives the predicted slope, and Equation 5-4 gives the slope determined by a linear fit of the data. (AR/R] = —6.1><10'°(Q-cmz)-l 5-3 tAp calc [AW—R) =-3.7x1093:.0.1x109(o.cm2)" 5-4 tAp exp 104 There is a 155% discrepancy between the predicted and measured values. This is similar to the findings of Hein and Schumacher,28 who reported a 72% discrepancy between the predicted value and their experimental results for O on Cu. It is a bit surprising that our discrepancy is even larger than theirs because their films very likely had poor surface conditions due to the fact that they were grown on spectroscopic glass slides and were therefore not epitaxial. The Persson scattering model is developed for a well ordered, smooth surface. As discussed in Chapter 1, Lin, Tobin, Dumas, Hirschmugl, and Williams” compared synchrotron measurements of AR/ R for O on a Cu(100) bulk single crystal with two different AR/ R predictions they calculated from two published tAp measurements on thin films. They found discrepancies of between the calculated and experimental values of 36% and 136%. The value that came closer to their measured AR/R was not far outside of the experimental error, and thus they considered the comparison to be encouraging for several reasons: (1) they compared a bulk single crystal AR/ R to a thin film tAp , (2) the thin films studied by others were poorly characterized, and (3) the wide variation in published tAp values indicated a poor degree of reproducibility of the film surfaces. Nevertheless, their large discrepancies are similar to that found in the careful simultaneous measurement of AR/ R and tAp presented here. ~ Here, in the first study of the linear relationship between AR/ R and tAp using simultaneous reflectance and resistivity measurements on a single, well characterized, epitaxial thin film sample, I find strong evidence in support of the linear relationship 105 predicted by the Persson scattering model. There is, however, a large discrepancy between the experimentally measured slope and that predicted by the model. 5-2. Formate Adsorption on an O-Predosed Cu(100) Thin Film The formate species consists of a hydrogen, a carbon, and two oxygen atoms ' (HCOO). It can be obtained through the direct decomposition of formic acid on Cu(100), but it has been shown that predosing the surface with O greatly enhances the sticking 29 We chose to examine formate adsorption on Cu(100) due to coefficient of formate. earlier indications that its broadband IR AR/ R behavior departs from that expected by the Persson scattering model. Lin, Tobin, and Dumas performed a reflection-absorption infrared spectroscopy (RAIRS) study of formate on a Cu(100) bulk single crystal in which they looked30 for a broadband AR/ R. They distinctly measured a broadband absorption when they predosed the Cu(100) surface with 0. Upon subsequent exposure to formic acid, however, the broadband AR/R returned to zero within the ~ 0.2% sensitivity of the synchrotron-based experiment. It was felt that a simultaneous measurement of tAp would provide useful information about the effect of formate on the scattering of conduction electrons in the metal. Figures 5-1 shows reflectance and resistivity time scans of formate adsorption on a Cu(100) thin film surface predosed with O. The size of the formic acid dose was 50 L, consisting of 5 x 10'7 Torr for 100 s at about the 500 5 point in the time scans. Figure 5-2 presents the data of Figure 5-1 in terms of AR/ R and Ap/ p , with linear drifi removed. It 106 can be clearly seen in Figure 5-2(a) that the broadband IR reflectance returns toward its clean surface value. There does appear to be a slight difference from AR/ R = 0, below the sensitivity level of the synchrotron-based RAIRS apparatus. The plot of Ap/ p reveals that the resistivity also returns in the direction of its clean surface value, although there is a distinct offset from Ap/ p = 0. The Persson scattering model deals with changes from a clean metal surface to a surface with an adsorbate, whereas our formate experiment measures the change from an O predosed surface. The reaction of the formic acid with the O replaces O adsorbates with formate adsorbates. The actual chemistry is not well understood. Nevertheless, a qualitative test of the linear relationship between AR/ R and tAp for formate adsorption on Cu(100) from an O predosed surface is interesting. Figure 5-4 is a plot of AR/ R versus tAp for the formic acid dosing region of the data shown in Figure 5-2. The data in Figure 5-4 comes from the 500 s to 598 3 part of Figure 5-2(a), and the 504 s to 602 8 part of Figure 5-2(b). These data regions were chosen to provide the most nearly linear plot. The plot in Figure 5-4 does not appear particularly linear, although it’s difficult to make a careful determination due to the fact that both AR/ R and tAp reach plateaus so quickly. The bulk of the data falls into the upper lefi hand corner of Figure 5—4, which is enlarged in the inset, where neither AR/ R nor tAp undergo significant change. A greater number of data points in the region of rapid change would allow a better synchronization of the AR/R and tAp time axes. 107 0.000 ... . ,., I 'fifir -0.002 _0004 p 0.0000 . . . d g 0.00041 1 00008: ; .. '1 O .0006 ~ 00012: ’° '. ,. .’ . -0.0016: .0.0020- ‘3 '0’008 " 1.06 A 1.08 A 1.10 ‘ I 1.0 1.2 A 1i4 A 11:6 4 ITS 2.0 2:2 A 2.4 tAp (106 flocmz) Figure 5-4. AR/R versus tAp for formate adsorbtion on a Cu(100) film. The inset is an enlargement of the upper left hand comer. The data are taken from about the 500 s to 620 8 region of Figures 5-2(a) and 5-2(b). As the formate replaces the O on the Cu(100) surface, the AR/ R and tAp induced by the 0 both return in the direction of their clean-surface values. This could be due to the removal of the O scatterers, provided the formate does not significantly induce conduction electron scattering. The fact that AR/ R appears to return very close to zero suggests that the formate does not scatter conduction electrons, but the fact that tAp does 108 not return nearly as close to zero suggests that formate does scatter conduction electrons. This discrepency, along with the apparent lack of a linear relationship between AR/ R and tAp , could indicate that the scattering model does not apply to formate adsorption on Cu(l 00). 109 References ‘K. Fuchs, Proc. Cambridge Phil. Soc. 34 (1938) 100. 2T. Holstein, Phys. Rev. 88 (1952) 1427. 3G.E.H. Reuter and EH. Sondheimer, Proc. Roy. Soc. London, Ser. A 195 (1948) 336. 411.13. Dingle, Physica 19 (1953) 311. 511.13. Dingle, Physica 19 (1953) 729. 6P. Wissmann in: Surface Physics, ed. G. Htihler, Springer Tracts in Modern Physics Vol. 77 (Springer, New York, 1975). 7D. Dayal, H.-U. Finzel and P. Wissmann in: Thin Metal Films and Gas Chemisorption, ed. P. Wissmann (Elsevier, Amsterdam, 1987). 8D. Schumacher, Surface Scattering Experiments with Conduction Electrons, ed. G. Htihler, Springer Tracts in Modern Physics Vol 128 (Springer, New York, 1993). 9M. Watanabe and A. Hiratuka, Surf. Sci. 86 (1979) 398. “’11. Merkt and P. Wissmann, z. Phys. Chem. Neue Folge 135 (1983) 227. ”M. Watanabe and P. Wissmann, Surf. Sci. 138 (1984) 95. 12J.E. Reutt, Y.J. Chabal and SB. Christman, Phys. Rev. B 38 (1988) 3112. 13OJ. Hirschmugl, G.P. Williams, F.M. Hoffmann and Y. J. Chabal, Phys. Rev. Lett. 65 (1990) 480. ”B.N.J. Persson, Phys. Rev. B 44 (1991) 3277. 15KC. Lin, R.G. Tobin, P. Dumas, C.J. Hirschmugl and GP. Williams, Phys. Rev. B 48 (1993) 2791. ”Charles Kittel, Introduction to Solid State Physics (John Wiley and Sons, Inc., New York, 1986) 144. 110 ”RC. Lin, R.G. Tobin, and P.Dumas, Phys. Rev. B 49 (1994) 17273. ”A. Spitzer and H. Luth, Surf. Sci. 118 (1981) 121. '9M.H. Mohamed and LL. Kesmodel, Surf. Sci. Lett. 185 (1987) L467. 20M. Wuttig, R. Franchy, and H. Ibach, Surf. Sci 213 (1989) 103. 2|H.C. Zeng, R.A. McFarlane, and K.A.R. Mitchell, Surf. Sci. Lett. 208 (1989) L7. 221K. Robinson, E. Vlieg, and s. Ferrer, Phys. Rev. B 42 (1990) 6954. 23MC. Asensio, M.J. Ashwin, A.L.D. Kilcoyne, D.P. Woodruff, A.W. Robinson, Th. Lindner, J .S. Somers, D.E. Ricken, and A.M. Bradshaw, Surf. Sci. 236 (1990) 1. 24Ch. Woll, R.J. Wilson, 3. Chiang, H.C. Zeng, and K.A.R. Mitchell, Phys. Rev B 42 (1990)]1926. 25R. Mayer, C.-S. Zhang, and K.G. Lynn, Phys. Rev B 33 (1986) 8899. 26NW. Ashcroft and ND. Merrnin, Solid State Physics (Holt, Rinehart, and Winston, New York, 1976) 5. ”WA. Reed and E. Fawcett, J. Appl. Phys. 35 (1964) 754. 28M. Hein and D. Schumacher, J. Phys. D: Appl. Phys. 28 (1995) 1937. 29M. Bowker and R.J. Madix, Surf. Sci. 102 (1981) 542. 3"RC. Lin, R.G. Tobin, P. Dumas, J. Vac. Sci. Technol. A 13 (1995) 1579. Chapter 6 Conclusions I have employed a novel form of reflection-absorption infrared spectroscopy, utilizing a unique liquid nitrogen-cooled grating spectrometer, to probe the effects of gas adsorption on the electron dynamics in the near-surface region of metals. Information gained from this study contributes to our fundamental understanding of surface properties. The results of this research provide support for a model due to Persson and Volokitinl'4 that relates a number of different physical phenomena, including a broadband absorption of infrared light and an increase in resistivity, to the scattering of conduction electrons by adsorbates. This work also provides impetus for further research regarding the scattering model. Due to the small size of the effect, the difficulty of maintaining a sufficiently stable baseline, and the low frequency range where nonlocal effects are significant, few measurements of the broadband reflectance change AR/ R have been reported?‘4 Here I present measurements of the adsorbate-induced change in the broadband IR reflectance AR/ R of Pt(1 l l) as a function of frequency, temperature, and CO coverage. The results are compared with the scattering model."4 The frequency dependence of AR/ R for CO on Pt(1 11) is at least qualitatively consistent with the theoretical prediction, but experimental nonreproducibility and uncertainty in the material parameters prevent a rigorous test. 111 112 The magnitude of AR/ R at 2500-2800 cm'1 is smaller at 90 K than at room temperature (~315 K) by about 30%. This result appears surprising, since the scattering model predicts that AR should be larger at the lower temperature.4 The clean surface reflectance R, however, is also larger at low temperature, with the result that a decrease in AR/ R at these frequencies is consistent with the theory. This demonstrates the previously unrealized fact that it is important to consider the temperature dependence of the clean surface reflectance R in addition to the temperature dependence of the absolute change in reflectance AR when dealing with the fractional change AR/ R. The magnitude of the reflectance change increases with increasing coverage for low coverages, peaks at a coverage of about 0.33 ML, and decreases toward saturation. The presence of a peak in the coverage dependence cannot be attributed to changes in adsorption site, but could be attributable either to partial ordering in the overlayer or to changes in CO's electronic structure. The slope of the curve at low coverage corresponds to a scattering cross section per CO on the order of 1 A2. The determination of this quantity is significant because it shows that broadband reflectance measurements can be used as a new probe to determine physically interesting quantities for systems where the scattering model is valid. Calculating and predicting the values and variations of these quantities, such as the cross section, presents a challenging and fertile field for theoretical investigation. It is significant that the Persson-Volokitin scattering model is successfill in treating the temperature- and coverage-dependence of CO on Pt(1 1 l), a transition metal surface. The scattering model assumes that the electrical conductivity is described by the 113 Drude free electron model. Since the Drude model is not expected to apply to transition metals, the success of the scattering model with regard to the CO/Pt(1 l 1) system provides valuable information regarding the extant of its applicability. Another prediction of the scattering model is that there exists a linear relationshipl between AR/ R and tAp , the adsorbate-induced change in resistivity (times the sample’s thickness). Previous research into this relationship was done either on dissimilar samples at different times” or simultaneously on polycrystalline films.16 I present here the first study of the predicted linear relationship using simultaneous reflectance and resistivity measurements on a single, well characterized, epitaxial thin film sample. The adsorption of O and formate on a Cu(100) thin film grown epitaxially by UHV evaporation onto an etched Si(100) substrate was studied. Strong evidence in support of the linear relationship predicted by the Persson scattering model is seen for 0 adsorption on Cu(100). There is, however, a large discrepancy between the experimentally measured slope and that predicted by the model. Linearity was not confirmed for formate adsorption on an O predosed Cu(100) film, but cannot be ruled out. As the formate replaces the O on the Cu(100) surface, AR/ R appears to return very close to zero suggesting that the formate does not scatter conduction electrons. tAp , however, does not return nearly as close to zero suggesting that formate does scatter conduction electrons. This discrepancy, along with the possible lack of a linear relationship between AR/ R and tAp for formate, could indicate that the scattering model does not apply to formate adsorption on Cu(100). 114 Taken together, these measurements provide support for the validity of the scattering model in explaining the reflectance changes induced by adsorption on metals, and indications of the limits of the model’s validity. In cases where scattering is the dominant effect of adsorbates on the electron dynamics of a metal, this work demonstrates that infrared reflectance measurements can offer a new probe of surface dynamics leading to accurate quantitative information. Interesting future work could include simultaneous reflectance and resistivity measurements of CO adsorption on Ni(100). It has been shown that an epitaxial Ni(100) ”1 film can be grown using a Cu(100) film as a “seed layer, 7 thus it is possible to produce a well characterized thin Ni(100) film appropriate for this type of study. The CO on Ni(100) system is interesting because an earlier test of the scattering model’s predicted quantitative relationship between AR/ R and tAp using AR/ R measured on a bulk single crystal and tAp measured on a thin film showed a discrepancy much larger than predicted8 — far larger than seen here for O on Cu(100). A second possibility is to further investigate the issue of coverage-dependent AR/ R. The adsorption of O on Ni(100) is a strong candidate system with which to test the possibility of an ordering effect. The O/Ni(100) system is known to have a single adsorption site, to adsorb in a disordered state at low coverage, and to undergo two ordering transitions observable by LEED at room temperature as the coverage is increased.18 Through these and other experiments, it should be possible to extend our understanding of the affect of adsorbates on the electron dynamics of metals. 115 References 1B.N.J. Persson, Phys. Rev. B 44 (1991) 3277. 2B.N.J. Persson, Chem. Phys. Lett. 197 (1992) 7. 3B.N.J. Persson and A.I. Volokitin, J. Electron Spectrosc. Relat. Phenom. 64/65 (1993) 23. 4B.N.J. Persson and A.I. Volokitin, Surf. Sci. 310 (1994) 314. 5D. E. Kuhl, K. C. Lin, C. Chung, J. s. Luo, H. Wang, and R. G. Tobin, Chem. Phys. 205 (1996) I, invited paper in special issue on “Surface Reaction Dynamics.” 61B. Reutt, Y.J. Chabal and SB. Christrnan, Phys. Rev. B 38 (1988) 3112. 7C.J. Hirschmugl, G.P. Williams, F.M. Hoffmann and Y. J. Chabal, Phys. Rev. Lett. 65 (1990) 480. 8RC. Lin, R.G. Tobin, P. Dumas, C.J. Hirschmugl and GP. Williams, Phys. Rev. B 48 (l993)2791. 9C.J. Hirschmugl, Y.J. Chabal, F.M. Hoffmann and GP. Williams, J. Vac. Sci. Technol. A 12 (1994) 2229. loK.C. Lin, R.G. Tobin and P. Dumas, Phys. Rev. B 49 (1994) 17 273 (1994); ibid. 50 (1994) 17760. ”CJ. Hirschmugl, G.P. Williams, B.N.J. Persson and A. I. Volokitin, Surf. Sci. 317 (1994) L1141. 12E. Borguet, J. Dvorak and H.L. Dai, SPIE Proceedings 2125 (1994) 12. '3 M. Hein and D. Schumacher, J. Phys. D: Appl. Phys. 28 (1995) 1937. l“CLA. Lamont, B.N.J. Persson and GP. Williams, Chem. Phys. Lett. 243 (1995) 429. lsK.C. Lin, R.G. Tobin, P. Dumas, C.J. Hirschmugl and GP. Williams, Phys. Rev. B 48 (1993) 2791. 116 16M. Hein and D. Schumacher, J. Phys. D: Appl. Phys. 28 (1995) 1937. l7Chin-An Chang, J. Vac. Sci. Technol. A 8 (1990) 3779. 18U. Starke, P.L. de Andres, D.K. Saldin, K. Heinz, and J.B. Pendry, Phys. Rev. B 38 (1988) 12277.