PLACE II RETURN BOX To roman this chockwtflom your "cord. TO AVOID FINESMumonorbdon ddocluo. I DATE DUE DATE DUE DATE DUE \ ” ‘ W i #J i ’— fiflfl MSU I. An mm W“ Oppomnny lm ADSORBATE-INDUCED BROADBAND INFRARED REFLECTANCE CHANGES ON METAL SURFACES By Keng-Ching Lin 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 1995 Hi br UR Sui re? 11“ ABSTRACT ADSORBATE-INDUCED INFRARED REFLECTANCE CHANGES ON METAL SURFACES By Keng-Ching Lin This research focuses on the broadband infrared reflectance changes of metal surfaces upon adsorption of foreign species. The 1-2% observed adsorbate-induced reflectance decrease is much larger than the 0.001% reflectance increase expected from the polarizability of the adsorbate itself. Therefore, this reflectance drop must result from changes in the dielectric properties of the substrate. This effect has received little attention since the signal is difficult to measure reliably, and most surface IR spectroscopy studies emphasize the vibrational modes of the adsorbate, which give rise to relative narrow absorption bands. By distinguishing the physical origin of these broadband changes, we can characterize the modification of electronic transport in the substrate by the adsorbate. This knowledge is important in understanding the fundamental physics of a wide range of technological subjects, such as mechanisms of chemical sensing and transport in thin films. I have done a series of experiments to investigate the broadband reflectance changes for different adsorption systems using two reflection- absorption infrared spectroscopy (RAIRS) systems, a synchrotron-based system at Brookhaven National Laboratory and a novel cryogenic spectrometer at Michigan State University. The experimental results provide a test of a proposed model of conduction-electron scattering from disordered adsorbates that incorporates nonlocal electrodynamics for the interactions of probing electric fields with conduction electrons. A direct relation between surface resistivity changes and reflectance changes is derived and discussed by comparing the measured reflectance changes and published resistivity changes on thin metallic films for CO on Cu, O on Cu, and CO on Ni. The results are consistent for CO and O on Cu. A discrepancy found for CO on Ni stimulates further discussion of the theory and experimental tests. For O on Cu(100) the reflectance change at different coverages were measured as a function of infrared frequency and fitted with a theoretical function. At coverages below 0.25 ML, the reflectance changes are well fitted by a universal form with a characteristic rolloff frequency that determines the transition between local and nonlocal optics. The coverage-dependence of the reflectance change is different from theoretical predictions. Possible explanations are investigated with reference to the adsorption systems of O on Cu(100), CO on Pt(111) and formate on oxygen-predosed Cu(100). To my Parents iv em ACKNOWLEDGMENTS The completion of this work is under numerous supports. I would like to express my deepest appreciation to my advisor Dr. Roger G. Tobin. Roger’ s intellect and insight of experimental physics provide me the best training I ever expected. His understanding of all the difficulties I encountered during these years encourages me to accomplish my best. Dr. Gwyn P. Williams designed the infrared beamline and Dr. Carol J. Hirsucmugl built up the surface science system at the National Synchrotron Light Source. Gwyn and Carol made this research possible by their primary works at the beamline. The collaborations with Dr. Fritz Hoffmann and Dr. Paul Dumas have been invaluable. Fritz and Paul’s scientific goals and personalities taught me a lot about being a researcher. Iwould like to thank all the members in Roger’ 3 group, Dr. Chilhee Chang, Dr. ].S. Luo, Dr. Hong Wang, Dennis Kuhl, E.T. Krastev, and Larry Voice. Chilhee handed me the rein of the system. Sam has always been a good friend. The discussions with Sam and Wang are very useful. Dennis came at the time when Ineeded a second hand to run the experiments and succeeds in expanding the project. The discussion with Ati and Larry of undergoing experiments is beneficial. I am also grateful to all my friends who V share my life in Michigan. They provide me joy, laughs and different prospects of life. My parents are my anchor for all these years. Being the youngest one of three children in my family, I indulge the love from my two sisters. The final thanks go to my husband, Woei-Wu Pai. His encouragement helps me to finish this work and head for a more challenge stage of life. vi List of ' list of i Chaple Chapte Chapte. FJ . ,._) Ix) I J [\J [\J . I\) Table of Contents List of Tables List of Figures Chapters Chapter 1 Introduction 1.1 Electron transport in a metal 1.2 Thin film resistivity studies 1.3 Surface RAIRS studies and surface effect on reflectance 1.4 Scattering theory of conduction electrons in adsorbed layer 1.5 Experimental designs References Chapter 2. Instrumental Introduction 2.1 Surface signal and light sources 2.2 Noise, detectors and relevant optics 2.3 Spectrometers 2.4 General surface science tools 2.5 Surface science branch at U4IR beamline, NSLS, BNL 2.6 RAIRS system at Physics Department, MSU 2.7 Summary and discussion vii vii viii 13 19 26 28 31 31 #88 53 56 60 Chapter 3. 3.1 3.2 3.3 3.4 Chapter 4. 4.1 4.2 4.3 4.4 4.5 4.6 4.7 Chapter 5. 5.1 5.2 5.3 5.3.1 5.3.2 References Adsorbate-Induced Changes in the Infrared Reflectance and Resistivity of Metals Experimental technique Surface effect on reflectance Comparison between reflectance and resistivity Summary References Adsorbate—Induced Changes in the Broadband Re- flectance of a Metal: Oxygen on Cu(100) Experimental procedures Data reduction Experimental results Frequency dependence Coverage dependence Dynamic dipole moment of oxygen on Cu(100) Summary References Infrared Spectroscopy of Formate on Oxygen-predosed Cu(100): Broadband Reflectance and Low-Frequency Vibrations Introduction Experiments Results and discussions Broadband reflectance change Copper-formate vibrations viii 62 65 72 78 80 80 86 89 95 100 102 103 106 106 108 109 112 5.4 Summary 115 References 117 Chapter 6. Adsorbate-Induced Reflectance Change of a Metal: 120 CO on Pt 6.1 Introduction 120 6.2 Experimental 122 6.3 Results and discussion: Coverage-dependence 128 6.4 Summary 137 References 138 Chapter 7 Conclusion . 141 Appendix 144 References 145 ix T. Table 2.1 Table 3.1 Table 3.2 Table 4.1 Table 6.1 List of Tables The performance numbers of two RAIRS systems at NSLS and MSU. Bulk material parameters for Cu and Ni. Adsorbate-induced changes AR,,/Rp in the reflectance of p-polarized light at the adsorbate coverage 11‘ as indicated, measured experimentally and calculated from Eq. 3.8 using the indicated values of tAp for thin films of thickness t, the electron density 1: given in Table 3.1, and an angle of incidence 0=85°. Best-fit parameter values and uncertainties obtained by fitting Eq. 4.2 to the spectra in Figure 4.2, and transforming the fitting parameters according to Eq. 4.3. The weighted average of the col values is 3383234 cm'l. Also listed are the values of a and a)1=co,,up/c found for CO on the same Cu(100) crystal, and the value of col predicted from a free electron model. Bulk material parameters for Pt. 61 73 74 92 128 ‘U . THC Wu .. . Wsfiv Fur Figure 1.1 Figure 1.2 Figure 1.3 Figure 1.4 Figure 1.5 Figure 1.6 Figure 1.7 Figure 2.1 Figure 2.2 List of Figures CO on Cu(100) surface. Distinguishable background absorption observed together with the dipole-selection forbidden hindered rotation mode of the CO molecule at 288 cm'1 and the Cu-CO stretch at 345 cm’1 over the frequency range 200-500 cm’l. A schematic of resistivity measurement on a metal film. A schematic of a deposited metal film on a substrate. CO induced resistivity change on a 100 A thick copper film as a function of coverage. Reflection of polarized light on a conductor surface. EP denotes the field parallel to the plane of incidence; Es denotes the field perpendicular to the plane of incidence. 0 is the angle of incidence and 8(a)) is the dielectric function of the substrate. A schematic of penetrating fields in the near surface region of a conductor. E 1 notes the field perpendicular to the surface; Ell notes the field parallel to the surface. The field inside the metal is attenuated. The near surface region of an adsorbate-covered surface. (a) The reflectivity of 2000 cm“ p-polarized light on Cu surface. (b) The reflectivity of 2000 cm”1 p-polarized light on Pt surface. The integrated vibrational signal on Pt surface with uncoupled oscillators model. The dashed line is for an oscillator with effective dipole moment and vibration frequency 2000 cm". The solid line is for an oscillator with effective dipole moment and vibration frequency 500 cm". xi 1O 13 14 15 22 33 33 figu. 1. , \hh.b . u . J. WU Pr. 3.. m. . WAN...» mLNw t . DU. 5313. U Figure 2.3 Figure 2.4 Figure 2.5 Figure 2.6 Figure 2.7 Figure 2.8 Figure 2.9 Figure 2.10 Figure 3.1 Figure 4.1 Figure 4.2 The radiation power from a blackbody source with (a) 0.5 mmzsr and (b) 0.05 mmzsr throughput, temperature 1000 K, 1 % efficiency and 1 cm“ bandwidth using i4 deg of the incident angle 84° and i3 deg of the incident angle 87° with a 25mmx5mm sample. Illustration of synchrotron radiation emission. Synchrotron radiation output curve with electron current 1 ampere, bending radius p 1.91 meters, 2 cm'1 bandwidth, OH 100 milliradians and summing over the vertical angle (iv. The bandpasses of selected optical components used in RAIRS experiments. Optical schematic of the Michelson interferometer. Czerny-Turner mounting of a reflecting grating system showing light at oblique incidence onto a grating and the definition of the angles 4) and 0. The schematic of the infrared beam extraction at synchrotron light source and the floor plan of surface science branch. Optical layout of the MSU RAIRS system. ARP/ Rp for clean Cu surface, CO/Cu(100), O/ Cu(100) and CO/Ni(100) over the frequency range from 300 to 2000 -1 C111. Ratio of the oxygen KLL to copper LMM Auger intensities as a function of oxygen dosage at 300 K. The coverage scale on the right axis is obtained by assuming a saturation coverage of 0.50 ML. Oxygen-induced fractional reflectance change AR/ R for four oxygen coverages: (a) 0.08 ML; (b) 0.17 ML; (c) 0.25 ML; (d) 0.35 ML. The gaps in the spectra and the noise at high frequencies are due to the low transmittance of the polyethylene windows. The solid lines are fits to Eq. 4.2 discussed in section 4.4; for spectrum ((1) no fit was possible. xii 37 39 39 43 47 55 57 67 83 88 L1 I (1‘; Pin CIT-4 - £3 :17 ,. (r? r. I 1:27;: 5v Scare Figure 4.3 Figure 4.4 Figure 4.5 Figure 5.1 Figure 5.2 Figure 6.1 Figure 6.2 Oxygen —induced reflectance change for (#025, with the instrumental offset C’-C subtracted. The solid line shows the best fit to Eq. 4.2 and the dashed line shows the results of Eq. 4.1 using the parameters given by Eq. 4.3. Over the frequency range studied, the two curves are virtually indistinguishable. Asymptotic (high frequency) limit of the reflectance change as a function of coverage (filled square). The values and uncertainties are taken from Table 4.1 for the three lowest coverages, and estimated from Figure 4.2 (d) for 0:0.35 ML. For comparison, the open circles show the resistivity change of a 40 nm-thick copper film. The low frequency spectrum from Figure 4.2 (c) for 0.25 ML oxygen on Cu(100) surface after removing the electron-scattering induced background reflectance change. The O-Cu stretch is observed around 340 cm". Adsorbate-induced fractional reflectance changes AR/ R due to (a) 0.25 ML oxygen on Cu(100); (b) formate on Cu(100) obtained by exposing the oxygen-predosed surface to 50 L formic acid. The data reduction procedures are discussed in the text. The solid line in (a) is the best fit to the electron scattering model. High resolution (6 cm'l) RAIR spectrum of formate on Cu(100), showing the bands assigned to the Cu-formate vibration. The lines show the best fit to a sum of two Lorentzians plus a linear baseline. Fractional change in reflected IR intensity from Pt(111) as a function of time at a frequency of 2800 cm‘1 and room temperature. The Pt(111) sample was exposed to CO for 100 seconds at about the 200th second. Background pressure during dosing was 8x10‘10 torr, while the pressure at the sample was increased by a factor of 16 due to the effusive array doser. A linear drift has been subtracted. (a) Fractional reflectance change AR/ R as a function of coverage for CO on Pt(111) at 315 K. Data shown were taken at two IR frequencies, 2500 and 2800 cm‘l. (b) Work function change for CO on Pt(111). xiii 93 97 101 111 114 125 130 Figure 6.3. The reflectance change lAR/R I vs. US, the ratio between the bulk electron mean free path and the skin depth, at a frequency w/ (01:4 where the reflectance change magnitude approach to an asymptotic value for the universal frequency-dependent function. xiv 136 mes tech: SON: inter; tells: infon the Spect LL18 b bells} of the hOW ; Slhch lTLil‘a f 3‘19er ‘3 mom Chapter 1 Introduction The interactions of adsorbed gases on metal surfaces are crucial to the fundamental understanding of heterogeneous catalysis and electrochemical reactions. The interplay between experimental studies and theoretical investigations along with the prosperous improvements of experimental techniques has led to a deeper understanding of the underlying physics of solids. Among the numerous experimental techniques that probe the interactions between substrates and adsorbates, light-based probes, such as reflection absorption infrared spectroscopy (RAIRS), have yielded fruitful information on the nature of bonding. The most common interest is to study the vibrational resonance of adsorbates on surfaces. However, the spectroscopic response of the substrate to light with various wavelengths, e.g. the broadband reflectivity, can also provide a clear picture of the interactions between adsorbate and substrate. Studying the changes in collective response of the substrate to the probing electric field upon gas adsorption can reveal how adsorbates modify the dielectric properties of metal substrates. In this work, I present experimental results obtained with a synchrotron-based RAIRS system, for the adsorbate-induced changes in infrared reflectance on metal surfaces and compare them to a scattering theory that incorporates nonlocal electrodynamics. The immediate motivation of this work comes from the measurement of the broadband reflec 1939 1 the d curl. Perssc the m. resuhs to V9: tramp.- region 2 reflectance change due to CO on Cu(100) by Hirschmugl and coworkers in 1989 [1]. They observed distinguishable background absorption together with the dipole-selection forbidden hindered rotation mode of the CO molecule and the Cu-CO stretch within the frequency range 200-500 cm'1 as shown in Fig. 1.1. The magnitude of reflectivity change increases monotonically with frequency and reaches an asymptotic value at high frequencies around 2000 cm'l. The significance of the broadband reflectance change is noted by Persson’s theory [2] which attributes the change of the dielectric properties of the metal substrate to the increase of diffuse scattering of the electrons that results from the damping of the parallel motion of adsorbates. It is important to verify the theory, which characterizes the effects of adsorbates on the transport properties of conduction electrons of metals in the near surface region. Electron transport in a metal will be reviewed in section 1.1 followed by a review of earlier studies of surface-scattering effects in thin film resistivity in section 1.2. A brief introduction about surface IR studies will be in section 1.3. Persson’s [2-7] theory of electron scattering and broadband reflectance change is introduced in section 1.4. A brief outline of conducted experiments is in section 1.5. a r - u - I - t _ c(2 x 2) CO/Cu(100] - -‘ 2 cm" _. E . l 5 2 O (I \ - '4 O O m 3‘. ’n - A J . n . on - 1 - 200 250 300 350 400 450 Frequency (cm") Figure 1.1 CO on Cu(100) surface. Distinguishable background absorption observed together with the dipole-selection forbidden hindered rotation mode of the CO molecule at 288 cm'1 and the Cu-CO stretch at 345 cm'1 over the frequency range 200-500 cm'l. ( reproduced from Ref. [27]) 1.1 the ma :pom matter's dielectr configu are treat level are field an Coulom and the ”collisic 879mm (0111310, be1/t metal Ca 1.1 Electron transport in a metal When electromagnetic waves are incident on a solid, the charges inside the matter rearrange themselves as a response to the external fields. The response of the matter to the electric field, e.g. the polarizability, defines the matter's macroscopic dielectric properties. A detailed understanding of the dielectric properties nevertheless calls for a microscopic picture of electronic configuration and electron transport inside the object. For metals, the ions are treated as fixed at lattice sites. The conducting electrons above the Fermi level are treated as a free electron gas which will be driven by the incoming B- field and cause the large polarizability of a metal. In the Drude model, the Coulomb interactions of a electron with other electrons and ions is neglected and the interaction which the electron experiences are only instantaneous "collision" events [8]. These collisions are assumed to maintain local thermodynamic equilibrium. If the mean time interval between two collisions is 1', the probability of collisions for a electron per unit time would be 1/ 1. Within the free electron model, the DC electrical conductivity of a metal can be expressed as: j= O'E , with 0' = p“1 = ne T Eq.1.1 m j : current density E : electric field strength 0': conductivity p : resistivity The me defined electron The Pet hmmt resStiri‘ model : r00m te electron mmdmd IQSlSfiVit the 0rde equafiOn mamfle This fem be...” my lOrm 5 n : electron density 1 : relaxation time 711 : effective electron mass e : elementary charge The mean distance an electron travels between two collision events can be defined as the electron mean free path : 1: DOT where no is the average electronic speed. For conducting electron, it could be the Fermi velocity vF . The Fermi velocity of the electron can be calculated with the known electron density: v; =(h/m)-(37tzn)2. The product of the mean free path and the resistivity depends only on the electron density within the free electron model : p-l=e‘2(37:2)” 311’“ 3. The order of the electron mean free path at room temperature is about hundreds of angstroms. For tungsten, the electron mean free path is about 144 A and 1140 A for copper which are hundred times of the interatomic spacing [9,10]. Room temperature resistivities are typically of the order of one microhm-centimeter which gives the order of the relaxation time about 10'14 sec for metals. The AC electrical conductivity of a metal can be given by solving the equation of motion for the momentum per electron and the current induced in a metal by a time-dependent electric field: 0'0 nezt _ . do = 1-zan' m Kw) = 0(w)E(w) ;0(w) = Eq.1.2 This form correctly reduces to the DC result at zero frequency. One point to bear in mind is that the equation of motion is derived by assuming a spatially uniform force. If the external field is from the propagating electromagnetic radiati of the compa: electric should in a m. Wavelei dielectr Where ( fiEldS ( frequm t“? of Path m, ,4 h I l 6 radiation in a metal, the validity of the Eq. 1.2 will be questioned. The effect of the magnetic field on the motion of electrons can always be neglected since the magnetic force is much smaller than the electric force. However, the current density at one certain point is entirely determined by the summation of the effect of the E-field on each electron at that position since its last collision. If the electric field does not vary appreciably over a distance comparable to the electron mean free path, the relation of current density and electric field in j(a),r)=0'(co,r)E(a),r) is justified. The wavelength of the field should be large compared to the electron mean free path about hundreds of A in a metal. This is satisfied by wavelength larger than visible light which wavelength is of the order of 103 to 104 A. Therefore the macroscopic complex dielectric function at infrared frequency region can be given by: 2 2 _ 47me2 p I a) = (02(1-1/ian) ” m s=1+i%g,s(m)=1+ Eq. 1.3 where (up is known as the plasma frequency. At low frequencies a)<>10), the fractional change of the resistivity can be expressed: r tip wil Q‘JAI‘Hllflt Ir material in thin ( and the 2 hard to < inner g1 EVidence mdepem Stages 0: material, rate, an Regimen forms, t} SubstTate 11 Ap 3 l0 - 1 . . —0 ——8( -p)—t Eq 16 As mentioned before the product of the resistivity and the mean free path is a constant depending only on the electron density 11. Since the electron density depends only weakly on the temperature and the defect density, the value of tAp will mainly depend on the effective specularity parameter which quantifies the properties of the film. In general, the properties of the metal film depend strongly on the material and deposition conditions. For example, recrystallization can occur in thin continuous films or during heat treatment, causing grain boundaries and the Fuchs specularity p of the same material can vary widely. It is usually hard to distinguish whether the scattering events of electrons happen at the inner grain boundaries or at the film surfaces. There is experimental evidence showing that the assumption of a homogeneous and thickness independent bulk-like structure is hard to meet in thinner films. The initial stages of film growth are strongly influenced by several factors such as material, structure, temperature and the purity of the substrate; evaporation rate, apparatus geometry, residual gas pressure and composition, sample treatment and measurement temperature [12]. Before a homogeneous film forms, the structure of the film is more or less like islands on top of the substrate. For complicai change 0: measurin; adsorptio: resistivity Wissrnanr ispnmari adSOI'ptior 100ml thic. on the 85 FtSlStii'jty the effect c Ap : where 3p i 12 For the study of the electrical resistivity upon gas adsorption, the complications introduced by film quality can be ignored if the fractional change of resistivity induced by the adsorbates is measured as long as the measuring conditions of the substrate remain the same before and after the adsorption. In the last two decades, the adsorption effect on the electric resistivity of thin metal films has been extensively investigated by Wissmann, Schumacher and other scientists [12,14-17]. The diffuse scattering is primarily responsible for the increase in the resistivity of thin films upon adsorption. Fig.1.3 shows the CO adsorption induced resistivity change on a 100 A thick copper film [12]. Although the detailed form of the curve depends on the gas species adsorbed, common features are a linear increase of resistivity change at low coverage and saturation at high coverage. To explain the effect on the resistivity of the gas adsorption, we can differentiate Eq.1.6: Ap Eq. 1.7 where Ap is proportional to the adsorbate density LA") =K'3fi K’a LA”) Eq.1.8 dna "rm f ' dna "‘40 This characterizes the initial slope of the measured curves of the resistivity change. The linearity between specularity change and the coverage is based on the assumption of the independence of scattering events. Adsorption can also affect the conduction properties by modifying the conducting electron density. Changes in n have been invoked to explain 13 thin-film resistivity changes due to surfaces, defects, and adsorbates, and oxygen incorporated at grain boundaries in a thin Cu film has been shown to reduce n [17]. But the influence of conduction electron density on surface resistivity and the reflectance of a single crystal surface remains undetermined. 50 . . a 40 ~ _ . J a O ' I I. g; 30 ~ 9' ~ E“. I' E 20 - a I,“ 10 - _ .. -' . I. - l u ' . o - ' i 4 I I . 0.0 0.5 1.0 1.5 n. (l 015 moleculdcmz) Figure 1.4 CO induced resistivity change on a 100 A thick copper film as a function of coverage. (reproduced from Ref. [12]) 1.3 Surface RAIRS studies and surface effect on reflectance Reflection absorption infrared spectroscopy has become a standard technique in surface science. The advantages of RAIRS include the high energy resolution, strict dipole selection rules, time resolution of dynamic processes, and convenient analysis of the vibrational lineshape and intensity. This technique has been successfully used to obtain useful information such 14 as the nature of chemical bonds of adsorbates, adsorption sites, molecular orientations, adsorbate-adsorbate interactions, and electrostatic effects of surface defects [18,19]. Among these investigations, the interest is usually focused on the vibrational absorption bands of the adsorbed species. Figure 1.5 shows a schematic view of the incident and reflected IR beams on a metal surface. interface-k ( ) W Figure 1.5 Reflection of polarized light on a conductor surface. denotes the field parallel to the plane of incidence; Es denotes the field perpendicular to the plane of incidence. 0 is the angle of incidence and 8(0)) is the dielectric function of the substrate. The s- and p- polarized electric fields of the radiation are perpendicular and parallel to the plane of incidence. On a metal surface, any dipole moment will induce an image dipole.Any molecular dynamic dipole parallel to the surface will be effectively screened out. For the dipole perpendicular to the 15 surface, the apparent dipole moment is enhanced. The vibrational modes with dynamic dipole perpendicular to the surface could be observed by p- polarized light through the electrodynamic interaction. When p-polarized light is incident at an angle 0 on a metal surface, the El field will decay more : electron mean free path 8 : classical skin depth Figure 1.6 A schematic of penetrating fields in the near surface region of a conductor. Ei notes the field perpendicular to the surface; Ell notes the field parallel to the surface. The field inside the metal is attenuated. rapidly into the surface than the parallel component E. . as sketched in Figure 1.6. Both components propagate exponentially into the metal with a characteristic skin depth 50 as shown in Fig. 1.6. The difference is that the value of 15.. at the surface is larger than by a factor ~\/|;|.The conducting electron in the near surface region will experience the external field and generate a current density corresponding to the field strength. This accounts 16 for the absorption of incoming light. The reflectivity of a metal will change, if the dielectric properties in the near surface region are modified. One can consider that the surface condition changes with a thin adlayer such as gas molecule adsorption or atom deposition. If the thin adlayer is a weakly- absorbing film (i.e. for a layer of oscillating dipoles, that condition will be satisfied far from the resonance frequency), the reflectivity change will be determined primarily by the changes of the optical properties of the substrate. Although people have been aware of the importance of this effect for the quantitative interpretation of RAIRS measurements [20,21], the background absorption caused by the modification of substrate conduction properties has usually been ignored. Standard analysis is based on the local response of the dielectric functions of metal substrates. The problem arises from the validity of the definition of the skin depth. If the incident light frequency a) is not too high i.e. much smaller than the plasma frequency, the classical skin depth is given by: 1/60 = c/Jeraa) where o is the bulk conductivity of the metal. The skin depth is around hundreds of A for normal metals [9,10]. The electric field is localized to a very thin layer at the metal surface. The validity of this picture is that the field in the metal varies little over the distance one conducting electron could travel during one cycle of the oscillation of the field. So it requires the frequency much larger thanvF/B which is defined as the characteristic frequency and the value is around 400 cm'1 [22]. When the frequent“) field ove: skin ef‘cct one cycle on reflect Nonlocali that an e motion 0 dielectric from a f'm estimate . Feibehnan llHIll of a meaSUre tl Parts repre this analt‘s POlaered 1 IEI-left a It Ce FOIE Scattenng 0] tr; . .' .EaQCllllt}. ‘ Bem'lEt [24] 17 frequency drops below vF/fi, the simple picture of an exponentially decaying field over a distance Sobreaks down. This phenomena is called anomalous skin efl'ect [8]. If the distance that an electron on Fermi surface travels during one cycle of field is larger than the skin depth, the calculation of surface effects on reflectance requires nonlocal corrections to the classical Fresnel equations. Nonlocality means a contribution to the time dependence of the electric field that an electron experiences during the time period 1/(1) comes from the motion of the electron in the spatially varying external electric field. The dielectric response at one certain point in space would be the summation from a finite spatial range defined by the travel of the electron. One can also estimate the dielectric properties in the near surface region using the Feibelman d-parameter formalism which is valid for C0>>Up/8 [23]. For the limit of a metal with no bulk absorption, the real parts of d-parameters measure the effective surface locations for the electric field and the imaginary parts represent the phase lag of the field in the near surface region. Within this analysis, the p—polarized light reflectance is far more sensitive than s- polarized light to the dielectric response. Even on a clean metal surface, the reflectance is so complicated that it is hard to get an analytic form. For good conductors such as Ag and Au, it has been shown that the scattering on the surface is specular if the surface is flat and free of defects. The reflectivity on these metal surfaces has been measured within local optics by Bennet [24] using the Drude model. The infrared absorption at a metal 18 surface would be larger if the collisions of the conducting electron with the surface are diffuse. The effective specularity of the conducting electrons can be used to characterize the reflectance change. Holstein used the specularity parameter to estimate the reflectance change mainly associating with the s- polarized field [25]. In order for surface-electron scattering to have a major effect on the reflectance of an Optically thick sample, two conditions must apply. The first is that the mean-free path 10 of the electrons must be at least comparable to the classical skin depth 80 of the metal. If 10 <<80 then the bulk scattering (electron-electron scattering and electron-phonon scattering) dominates the absorptance. The other condition is that the energy of the impinging radiation should be below the region where the interband transitions dominate the optical response of the metal. For good conductors at room temperature the range between 100 and 4000 cm'1 usually satisfies these two conditions [9]. The broadband absorption yields detailed information about the interaction between the surface and the probing E-field as well as how the adsorbate modified the surface. In the late-1980s, broadband reflectance changes upon gas adsorption were reported for H on W and M0 by Riffe, Hanssen, and Sievers [9] and Reutt, Chabal, and Christrnan [26]. Riffe and coworkers [9] observed a frequency-dependent reflectance change over the frequency range from 886 to 1088 cm'l. Qualitatively they attributed the changes of the attenuation coefficient of surface electromagnetic wave to the changes in the specularity of 19 scattered free carriers caused by the adsorbate-induced reconstruction of the W(100) surface atoms for high coverage adsorption. Riffe also studied other adsorbates and discussed free—carrier absorption. Later, Reutt and coworkers showed the reflectance change vs. frequency exhibiting a definite minimum for H on W and Mo which led to an interpretation of an intrinsic adsorbate- induced surface state around Fermi level. The possibility of the contributions by the free-carrier scattering was ruled out on the basis of a measured electronic band corresponding to the IR absorption. Hirschmugl et a1. [1,27] showed that the magnitude of the reflectivity change for CO on Cu increases monotonically with frequency and reaches an asymptotic value at high frequency region around 2000 cm'l. An asymmetric anti-absorption resonance assigned to the dipole-forbidden rotation of CO and one symmetric absorption of the Cu-CO vibration mode were also observed along with the broadband reflectance change. It is intriguing about the physical origins of the anti-absorption peak and the broadband reflectance change lineshape. The new theory proposed by Persson explains the underlying physical mechanism for the experimental results well. We will discuss the theory in the next section. 1.4. Scattering theory of conducting electrons in adsorbed layer The scattering theory of conducting electrons in an adsorbed layer, proposed by Persson and Volokitin [2-7], involves the damping of the parallel motions of the adsorbates on metal surfaces through electron-hole pair productit wmwn theory, th damping ; surface is expects tht because the screened or electronic c friction torn between the forbidden er electrons in line shape f‘ forbidden n decrease of affounts fOr Optics i5 am will Propaga “.cle of th e ( Elecmc field 20 production. It gives a microscopic picture of how the adsorbates modify the surface resistivity and the surface reflectivity. According the description of the theory, the diffuse scattering of the conducting electrons increases due to the damping process, hence the surface resistivity also increases. When the metal surface is probed by p-polarized light at near grazing angle incidence, one expects the surface parallel motions of adsorbates would not be detected because the dipole moment is perpendicular to the external electric field and screened out by the image dipole. If the dipole forbidden mode couples to the electronic current density induced in the metal by the external field, the friction force working on an adsorbate is proportional to the relative velocity between the adsorbate and the collective motion of the electrons. This dipole- forbidden excitation of adsorbate vibrational modes can be mediated by the electrons in the metal. In this case, an asymmetric band with anti-absorption line shape for the parallel vibrations is observed. In addition to the dipole- forbidden modes, the background broadband reflectance decreases. This decrease of reflectivity exhibits a frequency-dependent character which accounts for the local-nonlocal optics transition on metal surfaces. Local optics is accurate only if m>>vP/8. If 6<>n, the frequency is far away from the resonance of the parallel hindered motion of the adsorbate the reflectance (no, then the reflectance change is a term depending on the electron scattering rate: 4nM 1 1 cnmcosOrH Eq. 1.13 Then the direct relation between resistivity change and reflectance change is justified under the assumptions of the theory. lt try/5. It with cor optics. ' derivatic electric fi It An apprc relation t infinite 11 toward t] OSClllathr Which in a-‘Surned Can be Wr 5. Where T is EleClTOn C fiaCtion p ' Lhe YES] are llDl‘athn U 24 It is more complicated when the incident field frequency drops below vF/S. The problem is to figure out the induced current by the incoming field with conduction electrons colliding to the adsorbates under the nonlocal optics. The approach for the theory is briefly described below. A detailed derivation is given in Ref. [7]. For nonlocal optics, the spatially varying electric field have to be considered in calculating the induced current, [(x,a)) = [dax - o(x,x’;a))E(x’,a)) Eq. 1.15 An approach based on Boltzmann’s equation is used to obtain the non-local relation between current density and the electric field. If one assumes a semi- infinite metal with its surface in the x-y plane, the positive z-axis directed toward the interior of the metal, the adsorbate can perform harmonic oscillation in the x-direction, and also neglecting the normal field component which inside the metal is much weaker than the parallel component assumed along the x-axis, the electron distribution function f(z,vx,vy,vz) can be written in the form, .13: £511 in—f‘fo E 116 at+mau,+”‘az’ r q" where Tis the relaxation time and f0 is the Fermi distribution function. The electron distribution function will satisfy the boundary condition that a fraction p of the electrons arriving at the surface are scattered specularly while the rest are scattered diffusely with average no (the velocity of the adsorbate vibration in the x-direction), hmhi motion 0 the electr electron r. M: The electr nz; From Eq solution, calculated fer‘lectivih Per Scattering YEtlectanCE dellt'ed a l refleCtsince Mi \ R the 919C [IO 25 f(0,v,,vy,vz) = pf(0,vx,vy,—vz)+(1-p)f(vx —v0,vy,vz) Eq. 1.17 To take into account that a relative motion between an adsorbate and the drift motion of the electrons leads to friction clue to transfer of momentum from the electrons to the adsorbate, the friction force acting on an adsorbate is the electron momentum flux divided by the number of adsorbates per unit area: M55+Mm§x+ 11,, /n, = o Eq.1.18 The electron momentum flux in x—direction can be written as III: = m]d3vv,v,f(0,vx,vy,vz) Eq. 1.19 From Eq. 1.15-1.19 , the distribution function f(x,u) can be found. With the solution, the drift current induced by the external electric field Ex(z)can be calculated. Along with Maxwell equations, it is possible to calculate the reflectivity of the adsorbate-covered metal surface. Persson and Volokitin elaborate the nonlocal electrodynamics and the scattering effect of conduction electrons to calculate the adsorbate-induced reflectance changes on metal surfaces. The reflectivity for p-polarized light is derived and some numeric results are tabulated in Ref [7]. The fractional reflectance change can be expressed as follows: AR____—4a - (m/wl)2Re[g(a)/w1,l/6)] R 7:2 1_[—4—0)—]Im£; dq mop cosa 2(w/w1,l/5,q) Eq. 1.20 where g(x,z) and e(x,z,q) are complex-valued functions defined in Ref. [7],l is the electron mean free path, a is the magnitude of the reflectance change at 26 higher frequency and equal to %g—-—’;)vC—F, mp is the plasma frequency of the cos metal, 5=c/a)p is the classical skin depth, and (01 is a characteristic roll-off frequency, near and below which the nonlocal effects become important. Although the connection of electron scattering to vibraitonal damping and anti-absoprtion resonances is much emphasized in the theory, there are still questions in the assumptions [28]. If the frequency arise higher than the characteristic frequency (1),, Eq. 1.20 reverts to local optics and the reflectance change reaches an asymptotic value. The spectral shape of the broadband absorption depends on when and 1/6. All the quantities are properties of the substrate material. The interaction between the adsorbate and the substrate mainly determines the magnitude of the reflectance change, not the shape of the frequency dependence. The frequency dependence is the characteristic of broadband reflectance changes. The roll-off frequency characterizes the local-nonlocal transition and shows the conduction electron properties. The theory is based on a linear dependence on adsorbate coverages. It is of great interest to see how the adsorbate population would affect the reflectance change. 27 1.5 Experiment Designs My study is motivated by the proposed theory. As a test of whether the observed reflectance changes are related to the modification of electron transport, a comparison between resistivity changes and observed reflectance changes would show whether the scattering assumption holds in different experimental systems. A simple relation between measured resistivity changes on thin metallic films and observed reflectance change on a crystal is derived for the comparison. The relationship is confirmed within certain satisfaction for CO and oxygen on Cu surfaces. The details are discussed in Chapter 3. The frequency-dependent shape of the broadband reflectance changes , predicted by the theory is investigated with the system of atomic oxygen on Cu(100). The measured reflectance changes are fitted to the predicted functional form in chapter 4. At low coverages, the frequency dependence agrees well with the universal form. Some discrepancy remains at higher coverage. The characteristic of coverage dependence is also investigated. While the evidence suggests the direct interaction between adsorbate and substrate is through electron diffuse scattering, other adsorbate species show different results. I also study formate on oxygen predosed Cu(100) with the synchrotron based FTTR system and no reflectance change is found. Possible explanation is discussed in Chapter 5. 28 I use CO on Pt(111) as a further approach for the coverage-dependence. The nonmonotonic behavior of the reflectance change vs. coverage is similar to O on Cu(111). The reflectance change increases as coverage increases for coverages below 0.33 ML, reaches a peak at 0.33 ML, then declines toward the saturation coverage. The overlayer structure and work function change trend also switch at 0.33 ML. The correlation between reflectance change, overlayer structure, and surface electronic structure is discussed in Chapter 6. References: 1. CJ. Hirschmugl, G. P. Williams, F. M. Hoffmann and Y. I. Chabal, Phys. Rev. Lett. 65, 480 (1990). B. N. I. Persson, Phys. Rev. B44, 3277 (1991). B. N. I. Persson, Surf. Sci. 269/270, 103 (1992). B. N. I. Persson, Chem. Phys. Lett. 197, 7 (1992). B. N. ]. Persson, Phys. Rev. B 48, 15 (1993); B. N. J. Persson, J. Chem. Phys. 98(2), 1659 (1993). B. N. J. Persson and A. I. Volokitin, J. Electron Spectrosc. Relat. Phenom. 64/65, 23 (1993). B.N.]. Persson and A. I. Volokitin, Surf. Sci. 310, 314 (1994). N. W. Ashcroft and N. D. Mermin, Solid State Physics (Saunders College, Holt, Rinehart and Winston, Philadephia, 1976). 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 29 D. M. Riffe, L. M. Hanssen and A. J. Sievers, Phys. Rev. B 34, 692 (1986); D. M. Riffe, L. M. Hanssen and A. I. Sievers, Surf. Sci. 176, 679 (1986); D. M. Riffe and A. J. Sievers, Surf. Sci. 210, L215 (1989). KC. Lin, R.G. Tobin, P. Dumas, CJ. Hirschmugl and GP. Williams, Phys. Rev. B 48, 2791 (1992). K. Fuchs, Proc. Cambridge Phil. Soc. 34, 100 (1938). P. Wissmann, in Surface Physics, edited by G. Hohler, Springer Tracts in Modern Physics, Vol. (Springer, New York, 1975). EH. Sondheirner, Advan. Phys. 1, 1 (1952). D. Schumacher, Surface Scattering Experiments with Conducting Electrons, Springer Tracts in Modern Physics (Springer, New York, 1975). D. Dayal, H.-U Finzel, and P. Wissmann inThin Solid Films and Gas Chemisorption, edited by P. Wissmann (Elsevier, Amsterdam, 1987). H. Buck, R. Schmidt, and P. Wissmann edited by D. Hirschfeld (DGM_Informationsgellschaft, Wiesbaden, 1988), p.219. 1. Vancea and H. Hoffmann, Thin Solid Films 92, 219 (1982). YJ. Chabal, Surf. Sci. Reports 8 (1988); R.G. Tobin, Surf. Sci. 183, 226 (1987). FM. Hoffmann, Surf. Sci. Reports 3 (1983). R.G. Tobin, Phys. Rev. B 45, 12 110 (1992). R.G. Greenler, J. Chem. Lett. 44, 310 (1966); ].D.E. McIntyre and DE. Aspnes, Surf. Sci. 24, 417 (1971). 24. 26. 27. 28. 30 KC. Lin, R.G. Tobin and P. Dumas, Phys. Rev. B 49, 7273 (1994). PJ. Feibelrnann, Prog. Surf. Sci. 12, 287 (1982). HE. Bennet and ].M. Bennet in Optical Propertiesand Electronic Structure of Metals and Alloys, edited by F. Abeles (John Wiley and Sons, New York, 1966). T. Holstein, Phys. Rev. 88, 1427 (1952). LE. Reutt, Y. J. Chabal and SB. Christrnan, Phys. Rev. B 38, 3112 (1988). CJ. Hirschmugl, Y. I. Chabal, F. M. Hoffmann and GP. Williams, J. Vac. Sci. Technol. A 12, 2229 (1994). R.G. Tobin, Phys. Rev. B 48, 15 468 (1993). Chapter 2 Instrumental Introduction The major part of the experimental work was done at the U4IR beamline at the National Synchrotron Light Source at Brookhaven National Laboratory [1-5]. A home-built RAIRS system at Physics department, Michigan State University, is also used in this work [6]. These two systems are particularly designed for low energy studies, less than 2000 cm"1 (400 meV). I will discuss the aspects for the design of a RAIRS system in following sections, then introduce these systems at the end of the chapter. 2.1 Surface signal and light sources A generic RAIRS system consists of the source, spectrometer, UHV sample chamber, detector and relevant optics. Proper modulation can be chosen to separate a signal from superimposed background for sensitive detection. Primary concerns are related to optimizing the surface signals either due to the adsorbate vibrations or the changes in the surface optical properties of the metal. To optimize the performance of an experimental system, one seeks the highest signal and the lowest noise. The N/ S ratio value indicates the sensitivity of an operating system and expressed in term of root Hertz to take into account the measuring bandwidth [7]. The signal of interest in a RAIRS system is the fractional change in the intensity of the reflected light before and 31 32 after the gas adsorption on the substrate. This small change of the reflectance caused by the interaction of the substrate, the adsorbates, and the probing infrared radiation is obtained by comparing two spectral scans. Non- reproducibility of the raw spectra can appear as noise or drift in the comparison spectrum. High stability is therefore required for precise measurements. The absorption of incoming light by a solid surface is due to the polarization of the substrate which is related to the dielectric pr0perties and the dipole interaction on the surface. Gas adsorption can be modeled as a surface covered by a thin, homogeneous, and isotropic layer [8,9]. The surface region of a substrate where the dielectric properties are different from bulk values can also be treated as a distinct region [8]. If we work out the boundary conditions for the interfaces, the reflection intensity R (8, s) will be a function of the incident angle, and the dielectric properties of the layers. For the case of gas adsorption, the proper polarization can be chosen for the adsorbates which can link the microscopic parameters to the macroscopic dielectric functions [8]. The high conductivity of metals makes the reflection properties dramatically different for various polarizations of light. We are interested in p-polarized light due to surface sensitivity. Fig. 2.1 shows the incident angle dependence of the reflectivity on a bare metal surface using a Drude model for the dielectric functions of Pt and Cu [10] for p-polarized 33 1.0 . . . 0.9 3......“ .. >‘ 0.8 - ............. ‘ .2: . ' 3. «6 0.7 - .. .9. - . Q- o 0.6 _ ct a b i 0.5 - . ' 4 70 75 80 85 90 1.0 t .1 0.9 - . “uh-o... 0.00.0... ...... 0" _ .......... .. 3‘ ........... :E ..................... r 3 0.7 - ”m” -l 0 a . 0 i M 0,5 - ‘ OJ ' c1 70 75 80 85 9O Incident angle Figure 2.1 (a) The reflectivity of 2000 cm”1 p-polarized light on Cu surface. (b) The reflectivity of 2000 cm‘1 p-polarized light on Pt surface. Dielectric parameters are adapted from Ref. [9]. 34 light at the frequency 2000 cm'l. For p-polarized light at near grazing angle, 85-90 degree, less than 80% of incident beam is reflected. If the incident light is set close to grazing angle, the reflected background intensity is low and the fractional change due to the absorption of the adsorbates or the change of the dielectric properties in the near surface region is corresponding large. The grazing angle incidence is essential for detectability with the dipole selection rule for metal surfaces and the angle dependence 1/ c058 of the effective density of adsorbates. If one is interested in gas adsorption signals, the integrated fractional signals from the GO internal vibration and substrate- adsorbate vibration calculated using a non-interacting harmonic oscillator model on Pt are shown in Fig. 2.2. The absorptivity of the internal vibration is about several percent, while the external vibration is about half of tenth percent. So noise to signal ratio levels on the order of 10'2 percent are needed to get a convicing results for low frequency signals. The achieved sensitivities are 0.02% peak to peak Hz'”2 for the U4IR surface beamline [11] and 0.02% rms Hz'“2 for the MSU RAIRS system [12]. Both systems are marginally capable for adsorbate-substrate vibration measurements. However, the signal of interest in this study is broadband background reflectance changes which require the fluctuation of the absolute intensity of the source be less than the changes from surface effects about 1% to 0.2% over 20 mins. 35 0.4 ’ t ’ t ' l A E 3 _. 0.3 r ‘ a r"\ = ,’ \ .9.“ , VJ // F‘ I] ‘\ e x . .2 0.2 _ ’1’ 1—4 H ’ t o ” t a ,” | h r’ \. rt. . '0 ,o”’ '1 :5 0.1 v" if DD 1 0 l H l C t. H l/N l 000 . I i l ' l 1 7O 75 80 85 90 Incident Angle Figure 2.2 The integrated vibrational signal on Pt surface with uncoupled oscillators model. The dashed line is for an oscillator with vibration frequency 2000 cm". The solid line is for an oscillator with vibration frequency 500 cm". (reproduced form Ref. [6] ) 36 The existing apparatuses offer different advantages for broadband measurements. We will get back to the comparison later. For now, we just keep the typical signal of interest in mind. When one sets the incident angle close to grazing for optimizing the surface effect to the signal, the amount of radiation reflected off the sample will drop sharply if the solid angle of reflection at the sample becomes smaller. For an optical system, the acceptance of light for the apparatus can be defined as throughput: Throughput = accepting area xthe solid angle of acceptance 2 AQeffl If the incident angle approaches grazing incidence, the solid angle accepted by the sample will be reduced. One should keep a certain level of radiation to minimize the contribution from the source fluctuation to system noise. The typical size of single crystal for surface measurements is less than 2 cm2. The trade-off between maximizing the surface signal and throughput is important in designing a RAIRS system. In most cases, the sample is the optical component to limit the thoughput for grazing incidence and practical size. Fig. 2.3 shows the output spectra expected from a conventional 1000K globar IR source limited by the different throughput of a typical RAIRS system. 1In optics, people usually use the f-number(f/#) to characterize the ”speed" or the angular acceptance of an optical device. The f/# is equal to the ratio of the effective focal length of an optical component and the diameter of the circular device. The photon flux density varies inversely to the square of the f/#. 37 l 0'7 IIIIIIP ' l l 0-8 ITIIHII[" 'l l 0‘9 I rrrrrr] *1 \ i 1 1 11111 Watts \ 1 10-10 r rrrrrrr'r'r 1 1111111-.. L1 ij' 10‘11 \ 4111 n 1111111 -1 10-12’ 11'111 It"'rl . ..... 10‘ 102 103 Wave number (cm-1) Figure 2.3 The radiation power from a blackbody source with (a) 0.5 mmzsr and (b) 0.05 mmzsr throughput, temperature 1000 K, 1 % efficiency and 1 cm'1 bandwidth using :4 deg of the incident angle 84° and :3 deg of the incident angle 87° with a 25mmx5mm sample.2 [5]. 2The power from a thermal source with temperature T at frequency v in the unit of wave number , equal to 10., in an interval Av for a particular polarization is given by the Planck distribution law: pvdv = hc2v[exp(hcv / kT) — 1]". The throughput is calculated as : sample size x cos(incident angle) x sqr(expanding angle) 38 So a brighter source such as synchrotron radiation may be a solution to this restriction since the radiation is emitted into a cone with high flux density. The emission from an accelerated electron is shown in Fig. 2.4. The total emitted power emitted per electron P(o), Q, u) is a fuction of frequency, the solid angle, and the electron velocity from classical electrodynamics. The emitted radiation from a synchrotron light source can be expressed in practical units, per horizontal milliradian per ampere into a 0.1% bandwidth with electrons in the storage ring, integrated over vertical angle 8v as [13]: le=2.547x1013xsxcl(y) where E is the electron energy in GeV, G1(y) is the Bessel function with y=Ac/h and Ac is the critical wavelength which bisects the power spectrum and is given by k(A)=5.59xp/E3 where p is the bending radius of the storage ring in meters. The universal synchrotron radiation output curve is illustrated in Fig. 2.5. For the NSLS VUV ring at Brookhaven at a wavelength of 10 um (1000 cm'l), operating at 0.75 GeV, bending radius 1.91 meters, electron current 1 ampere, collecting 100 milliradians horizontally (about 6°), 1 milliradian verticaly and assuming 2 cm'1 bandpass, that corresponds to F.(,.=9.9x10'8 Watts [5]. For a 1000K black body soure at 1000 cm' 1 with 1cm'l bandpass, an opening throughput 0.5 mmzsr (typical throughput for RAIRS measurements), and 10% emittance the power is 1.8x10'8 Watts. These numbers are close to real experimental situation. 39 electron current cross section radiation area Figure 2.4 Illustration of synchrotron radiation emission. (reproduced from Ref. [5], p.23) .- J-I I . I - I .17 ~ I I I 1 10-6 :- I I :4 E I I 2: I ' ' 1' I. l a I T I I '1 I I I 1 10-7 ;- I I 1 I I 1‘ I ' ' I .- I I 4 I”, .. I I . z __ I I ‘ . I . . 3 E Infrared reglon . 1 Ultra Violet region 1 10~8 e . . 1 ': I I : '_’ I I I“ I- I I at >- I I -I I I . . o .4 g . .4— VISIble hght ] . I I 10’9 :' I I 1 : I I .1 : I I 1 .. I I h- I I ~1Hl 1-1 1 1111111-1 I L'II‘JII ' 1;;1-141 - I -.r,.-I . 1'I' 103 104 105 105 107 108 Wave number (cm-1) Figure 2.5 Synchrotron radiation output curve with electron current 1 ampere, bending radius p 1.91 meters, 2 cm'1 bandwidth, 8H 100 milliradians and summing over the vertical angle 8v. (reproduced and rescaled form Ref. [5]) 40 A synchrotron source is brighter than a blackbody radiation source. However, the synchrotron source intensity is proportional to the electron current in the storage ring and strongly depends on the geometry of the electron orbit. The source fluctuation turns out to be the major uncertainty of the experiments. The experiments were done with only limited adjustments available to offset the synchrotron source fluctuations. A strict data selection criterion is essential for data taken at NSLS. It will be discussed in chapter 4. 2.2 Noise , detectors and relevant Optics The noise we are concerned with is related to the detected signal. It is important to characterize the intrinsic noise from the detector and the noise contained in the signal arriving at the detector. There are several possible sources of noise from the system: source intensity fluctuation, ambient photon noise, mechanical instability, electrical noise, and the detector noise. Mechanical vibrations usually result in low frequency oscillations which can be chopped out by high frequency modulation [6]. Photon shot noise either from the source or from the ambient cannot be eliminated by lock-in techniques. The high ambient photon flux can saturate sensitive detectors. A cooled optical system can reduce the background radiation [6,7]. The designs of a cooling spectrometer and relevant optics take into account of cryo-state properties, such as thermal contraction which would affect the optical alignment, lubrication for mechanically driven systems, etc. [6]. One major 41 advantage of brighter sources is that larger ratios of signals to noise can be obtained with less effort to control the ambient thermal flux. The noise—equivalent power NEP (or the noise-equivalent photon rate NE N) is defined as the amount of radiant power (photon rate) that must be incident on a detector to produce a signal equal to the RMS output noise of the detector in a one Hz bandwidth. The NEP(NE N) is thus measured in W/ Hz”2 (photons/secHzm). The noise level of the detector indicates the sensitivity. Typical RAIRS requires NEP better than 10'10 W /Hz“2 to achieve the desired sensitivity of the system [4] for a black body source, less than 1%. Modern photoconductive detectors operated at low temperature have noise levels close to the photon rate 107 per sec, corresponding to the NEP=2X10'17 W/Hz“2 at 1000 cm’1 with conventional amplifiers [7]. The photoconductive detector generates a current proportional to the incident photon flux and produces a voltage across a load resistor. Specialized amplifiers are available for noise levels below ~100 photons/ sec. Photoconductive detectors have a cut-off frequency due to the impurity binding energy of the material which limits the probing frequency region. For low frequency broadband measurements, thermal detectors can be a better choice. The most sensitive thermal detectors are semiconductor bolometers. In low backgrounds, such bolometers have NEPs approaching 10'15 W/ Hz”2 when operated at liquid 4He temperature [7]. The resistance of the detector element will change exponentially with the temperature rise responding to the incident power 42 when a constant current is flowing in. A voltage proportional to the signal which is generated across a load resistor can be recorded by the data acquisition system. The measurable frequency range is an important element for a RAIRS system. The spectral range is determined by transmission of the optical components such as windows, the beamsplitter of interferometer, filters for the grating spectometer, and the operating range of the detector. A sketch for the band passes of selected windows, the beamsplitters, and detectors is shown Fig. 2.6. One should be aware of the Fabry-Perot interferences caused by the windows which produce fringes in a period of 1/ dn on output spectrum where d is the thickness in centimeter and n is the refractive index of the window material. For example, a Csl window with thickness about 7 m m will result in the fringes ~ 0.4 cm'l, which won’t affect the spectrum with resolution 1 cm'l. 2.3 Spectrometer A chosen spectrometer will give the required resolution of the measured spectrum and should not limit the throughput. Two types of spectrometers are used in this study. 43 l ' r ' l diamond window silicon window/beamsplitter (dashed line transmittance < 10% ) .------------‘----------------‘-------- cesium iodide window KRS-S polarizor substrate graded polyethelene (dashed line, major absorption bands) ......................................... Cu doped Ge photoconductive detector Silicon bolometer 0 500 1000 1500 2000 Frequency (cm'l) Figure 2.6 The bandpasses of selected optical components used in RAIRS experiments. ( modified from Ref. [5]) 44 A. Fourier Transfrom Infrared Spectrometer (FTIR) A Fourier Transform Infrared Spectrometer (FTIR) consists of a Michelson interferometer, a fast analog-to—digital converter, and a computer station with a fast Fourier Transform calculation algorithm. The schematic of a Michelson interferometer is shown in Fig. 2.7 [5]. The collirnated source input is divided by the beamsplitter into two components. One is reflected back by the fixed mirror, the other by the moving mirror. They are recombined at the beamsplitter and directed into the detector. The signal at the detector is a function of the optical path difference between the fixed and moving mirrors that results in an interference pattern. The interferogram can be Fourier-analyzed to reconstruct the broadband spectrum as a function of frequency [14]. fixed mirror It D ‘ n I beam splitter movmg mirror 4— _> detector D All ‘ 4 ¥ 1 L/ 2 light source Figure 2.7 Optical schematic of the Michelson interferometer. (from Ref. [5]) 45 Three major advantages over dispersive monochrometers for a FTIRS, the throughput advantage, the multiplex advantage and the rapid-scan mode providing a modulation of the signal [5,14]. The resolution of an interferometer is determined by the total path difference of the two interfering rays. The theoretical maximum value the resolving power is: RaZd Mo, where 110 is the wavelength of the incident monochromatic radiation and d is the maximum optical path difference. The resolution doesn’t depend on the aperture size. When one chooses higher resolutions, the throughput will not be degraded. This is the throughput advantage; it is not necessary to reduce throughput to get higher resolution. The multiplex advantage (Fellgett’s advantage) exist if the noise is independent of the spectral bandwidth over which signal power falls on the detector, such as the inherent detector noise. The FTTRS can measure the entire spectrum of N spectral elements in the same time that a grating spectrometer with similar efficiency and throughput takes to obtain a single spectral element with the same signal to noise ratio. An improvement of a factor of N”2 is gained compared in the normal situation. The signal sampling can be triggered by a He-Ne laser that shares the optical path. The moving mirror is traveling at a constant velocity and the signal is recorded by the fast AD converter sampling at fixed path difference 46 intervals. This mode provides a specific modulation frequency, which is proportional to the mirror velocity, for every incident wavelength. To study the frequency dependence of the adsorbate-induced broadband reflectance change, the convenience of getting the whole spectrum range in one scan is the major reason one chooses a FTIR. However, although FTIR also offers various advantages, for RAIRS, sample limits throughput and high resolution isn’t necessary— surface vibration modes are usually wide. The measurements are never really limited by noise, but by signal fluctuation or instabilities—multiplex advantage may not exist either. The simplicity and ability to do monochromatic measurements with a grating spectrmeter can give better sensitivity to changes in broadband reflectance. B. Reflection grating spectrometer The Czerny-Turner mounting shown in Fig. 2.8 of a reflecting grating is widely chosen for reflection grating spectrometer [6]. The resolution of the spectrometer is determined by the dispersive power of the grating, the slit width and the focal length of the focusing mirror. The throughput is determined by the slit size and the f/# of the collimator. A larger slit width will give higher throughput, but degrade the resolution. It is crucial to maintain adequate resolution and maximize the throughput to increase the sensitivity. The angular resolution of the spectrometer is determined by the slit width w and the effective focal length f of the focusing mirror: d8=w / 2f. 47 collirnating mirror \ optical axis fightout lightin _ e gran' ; normal _— — — _— _— — — — — — — _— __- .— — Izgrating anJe grating focal point focal point entrance slit exit slit Figure 2.8 Czerny-Turner mounting of a reflecting grating system showing light at oblique incidence onto a grating and the definition of the angles o and 8. 48 The spectral resolution is also related to the dispersion of the grating. A grating spectrometer has to be calibrated to get accurate frequency measurements. The calibration is done by using a known monochromatic source to get the spectra of different order diffraction peaks compared to calculation. A He-Ne laser can be used. The diffracted angle of the beam line of n-th order is expressed as the grating equation: n}. = 2dsin0n cos¢ where A is the wavelength of the laser, 6328A, n the order of diffraction, d the grove spacing of the grating, 6,, the grating angle at the n—th order peak, and 4) the half angle between the incident and diffracted beams of light on the grating face as shown in Fig. 2.8. By measuring the laser spectra of several orders along with 0“, cost) can be determined from the slope. By differentiating the equation and dividing them, one can get the spectral resolution and resolving power: (12. dv fix 1 A v 2f tan0 In the system in MSU, the grating spectrometer is cooled to reduce the background radiation [6]. Measurements show different coso values for room temperature and liquid nitrogen tempertaure; 0.983 and 0.980. It is consistent with the thermal contraction of aluminum which is the substrate material for the grooved grating. Filters are used to remove the higher than first order 49 diffraction which would contribute about 10% to the signal if it were not removed. 2.4 General surface science tools A contamination-free metallic surface has to be maintained for the study of the interaction between adsorbate and metal substrate. A clean surface is prepared under well controlled conditions to ensure the absence of foreign matter. This requires the development of an ultrahigh vacuum chamber to keep clean for a period of time, cleaning tools and recipes inside the UHV, and also reliable experimental techniques for characterizing the solid surface. A working UHV system should maintain the pressure on the order of 1x10'10 torr. That leaves the sample free of the gas adsorption about one monolayer for 10,000 secs [15]. The UHV techniques primarily concern the pumping system designs, from roughing pumps for low vacuum to ion pumps or turbomolecular pumps for ultrahigh vacuum. In order to reach ultrahigh vacuum, a bake- out process of the metal chamber wall is required to reduce the outgassing rate at room temperature. In general, stainless steel is chosen as the material of the chamber for the strength to the pressure difference and the low out- gassing rate at high vacuum. High quality aluminum alloy can be used sometimes. The titanium sublimation pump is usually equipped for removing residual gas molecules after extensive gas dosing in the chamber. 50 The main chamber has to be sealed below atomic leaking level. Knife edge flanges with soft metal gaskets are used for UHV. Technical information is available with vacuum product manufacture’s catalogs [16]. Proper surface analysis tools can reveal useful informations for a clean surface while UHV provides the suitable environment. Standard surface analysis tools are including: Auger Electron Spectroscopy (AES), Low Energy Electron Diffraction (LEED), and Temperature Programmed Desorption (TPD). Auger electron spectroscopy is widely used to determine the chemical composition of a clean surface. A high energy 2-5 eV electron beam is directed on the sample and the backscattered electrons of a selected energy N(E) are collected. A typical Auger process is: the incident electron collides with an atom and ionizes a ls electron, a 25 electron drops into the hole and the transition energy ejects a second electron, Auger electron from the 2p level. This Auger process is called KLL3 . To distinguish Auger electrons with characteristic energy from scattered electrons which lose their energies in the inelastic collisions, the derivative signal dN(E)/dE is more sensitive to identify the peaks from Auger transitions. The electrons with kinetic energies in the range 15-1000eV have a very short mean free path in matter (<10 A). 50 the signals are from the top layers of the solid and the information obtained is 3The atomic levels in an Auger transition are labelled in accordance with conventional X-ray spectroscopic nomenclature, i.e. K, L, M, for the n=1, 2, 3,... principal quantum numbers of the atimic shells. 51 about surface properties. The binding energy of a core level electron is a sensitive function of atomic species. Auger electron energies are well known and tabulated [17]. Every element with Z > 3 exhibits some Auger decay within the critical range for surface sensitivity. Since every surface atom leaves its fingerprint in the kinetic energy spectrum, Auger spectroscopy is suited perfectly for surface elemental analysis. Using AES data for quantitative analysis requires a well-calibrated reference system. Otherwise the information will be limited to identify the contamination species on surfaces. The sensitivity can usually reach about 1%, i.e., the impurity density higher than 0.01 ML will show up in Auger measurements. Low energy electron diffraction is used to observe the surface structure. The LEED pattern is an image of the surface reciprocal net when viewed along the surface normal at a great distance from the crystal. The existence of a sharp spot pattern implies the existence of a well-ordered surface and provides direct information about the symmetry of the substrate. Translational invariance in two dimensions ensures the diffraction occurs if the two-dimensional Laue conditions are satisfied: (k. —k,)-as =2nm; (k. —k,)-bs =27m, where ki and k; are the wave vectors of the incident and scattered electron, a5 and b3 are the primitive vectors defining a unit mesh, m and n are integers [15]. If adsorbates form an ordered structure on top of the substrate, the expected pattern is found by superimposing the diffraction pattern induced by 52 the adsorbate-overlayer and the substrate diffraction. In principal, one can decide the overlayer coverage by an observed LEED pattern and use the information for coverage calibration for ABS and TPD. After adsorption happens on a surface, desorption becomes one of the elementary surface processes. The desorption rate of adsorbed species on a substrate depends on the desorption energy and the temperature. If one ramps the temperature of the substrate in a controlled way, the desorption will happen and a measurable quantity, e.g. partial pressure of the adsorbed species, can be related to this process. By examining the temperature profile of the desorption, one can either get the information of the desorption energy [15,18] or integrate the desorption curve to obtain relative coverage informations for adsorbates. However, quantitative measurements also require well-calibrated reference system and residual gas analyzers which are usually subjected to change to different vacuum conditions. Contaminations on the surface can be removed by heating or ion sputtering. Sputtering process is performed in low inert gas pressure around 1><10'4 torr with an electron gun to ionize the gas. The ions collide on surface and remove top layer atoms. In general, after sputtering the surface becomes rough and annealing is required to restore a atomic flat sample. Gases are leaked into the UHV chamber through metal seal leak valves by backfilling the chamber or with specilly designed dosers which enhance the dosage on sample surfaces [20]. 53 2.5 Surface sciencebranch at U4IR beamline, BNL The basic optical layout of the RAIRS system at the BNL, U4IR beamline is sketched in Fig. 2.9 [2,3,5] The source radiation is extracted from the electron storage ring by mirror M1, travels through transfer optics, and is introduced to the surface science branch platform after M5 mirror. The opening angle of the beamline is 90 mrad x90 mrad to preserve the source brightness at long wavelength. M1 is a silicon-carbide mirror coated with a highly conducting surface, also equipped with water-cooling system to dissipate the heat generated by the radiation. The ”periscope”, M1 and M2, sends the beam in the upstream direction to the ellipsoidal M3 which deflects the beam horizontally by 90°. M4 deflects the beam vertically to another ellipsoidal M5. An image of the source with a magnification 0.9 is produced between M4 and MS with a slit assembly for alignment. The ring vacuum of ~10'10 torr is separated from the spectrometer vacuum ~10'3 torr by a piece of type 2A diamond window with thickness 0.8mm which can transmit the whole IR region and visible light for alignment. But it produces Fabry-Perot interference fringes. Measurements with resolution higher than 4 cm"1 cannot be achieved. It is later modified with a CsI window for high resolution measurements. After the diamond window, the diverging synchrotron beam is collirnated and directed into the spectrometer. The collimation is achieved by a combination of a spherical mirror and a plane mirror. 54 A modified Nocolet 20f Michelson vacuum interferometer is used for both transmission and reflection measurements. After the interferometer, the collimated light is focused onto the single crystal sample in the UHV chamber with f/ 10 optics at 85° angle of incidence. The reflected beam is collected by an off-axis f/3.5 ellipsoid and directed into the detector. The sample manipulator is facilitated for the translation motion along x-, y-, and z-directions and the 360° rotation to reach AES, LEED, sputtering and infrared positions. The material for the windows on the UHV chamber in the optical path was polyethylene which permits 80% of light in the range of 10-600 cm'l, but has strong absorption bands in the mid-IR (600-2000 cm'l). Two detectors were used to cover the frequency range from 80-2000 cm'lz a boron-doped silicon bolometer at 4.2 K for low frequency measurements 80 to 650 cm'1 and a copper-doped germanium photoconductive detector also at 4.2 K for higher frequency measurements 300 to 2000 cm'l. The thermal equilibrium state of the detector is crucial for the experiments. The bolometer is equipped with liquid helium cooled filters to reduce the high frequency radiation background. There is a Winston cone in front of the detector element which collects the incident beams within a fixed solid angle as a converging lens. 55 UW Chamber NSLS SOURCE Kinematic Mirror —- I (Roll-cud Upwardl From Below) \ Nlcale! 20F Michelson ,Ptom Mina lnlulcromglu Otomond Window Elliplicol Minor M5 to experimental branch M2 / / M4 \ bending radius 1\ electron beam / I 4’ / \ infrared radiation 33.5 deg M1 port Figure 2.9 The schematic of the infrared beam extraction at synchrotron light source and the floor plan of surface science branch. (modified from Ref. [3,4]) 56 2.6 RAIRS in physics department, MSU The basic optical layout of the RAIRS system in MSU is shown in Fig. 2.10 [6]. The source is a conventional silicon carbide globar which is surrounded by a liquid-nitrogen-cooled housing. The globar is heated resistively to approximately 1300K by an HP model 6264 B power supply. The source cryostat is pumped out by an Ultek ion pump to about 1x10° torr and cooled by liquid nitrogen. An automatic liquid nitrogen feeding device is installed. The LN2 can is supported by three stainless steel rods which are kinematically mounted on the bottom of the source chamber. The radiation is collected, reflected, and focused by an 60° off-axis ellipsoidal mirror to an exit slit where the tuning fork chopper is located. A plane mirror is also mounted on a turret mirror mount for reflection of a laser beam coming though a window from outside of the cryostat into the infrared optic axis. It is usually out of the infrared optical path and is moved in for alignment. A tuning fork chopper is chosen for the operation in vacuum and at liquid nitrogen temperature. The tuning fork chopper is a custom made one from Multi-scanning System Corp. type RC-2 operating at 800 Hz. The assembly consists of the fork, a pick up coil, a driving coil, a magnet and a mounting base. The AC voltage generated by the motion of the vanes in the pick-up coil is amplified and the amplified voltage is applied to the drive coil in the proper phase to cause the vanes to vibrate at the fork’s resonance frequency. 57 l! ran-mares mace mecrmm msrcu norm. ~Cooled ! Spectral rmqe: :00 - woo a-l Grating Spec router Resolution: I- 5 a—l Saple lap: 20 - 1200 K ._.. . . . . . eon-om coo-_- Ultrdridr Vocqu Sirloce Analysis Outer uith Single Crystal Saple on Ute-cooled uripulotor Uic~Cooled Si :8 lnira’ed Detector Other striooe probes: ll! Source 0703“! Low [terry Electron Diffraction IIUI LII-cooled SIIeIds. W [lectrm hum ‘ x 50 I), [ll Polmzer and Clara-er "ml Desorption Spectroscopy Figure 2.10 Optical layout of the MSU RAIRS system. The source, chopper aperture, sample, spectrometer entrance and exit slit are conjugate points. (from Ref. [6]) 58 The signal from the amplifier is synchronous with the vane motion and used for a reference for modulation. All the optical components are mounted on a thick copper plate which forms the bottom of the liquid nitrogen can. the cooling of the optical components can be monitored through Si diodes connected with electrical feedthroughs. There are two transfer optics; one in between the source cryostat and the UHV chamber, and the other in between the UHV chamber and the spectrometer. Together they enable the infrared radiation to illuminate the whole sample with an adequate size of the infrared windows and to be collected by a spectrometer with a relatively low focal number f/4.2. The UHV chamber is a custom-built stainless steel ultrahigh vacuum UHV chamber 12" o x 25” h. The chamber is divided into three levels: infrared level in the lowest, surface analysis level above the IR level and a manipulator on the top. The travel of the manipulator can be 1” in x- and y- directions and 10” in z-direction and rotated 360° with a differentially pumped rotary seal. Surface analysis tools are equipped on the surface analysis level. A Perkin-Elmer Model 10-155 Cylindrical-Mirror Analyzer is installed for AES analysis. A reverse view low energy electron diffraction optics (Princeton Research Instrument Co.) is used to investigate the surface structures. A Varian Ion Bombardment Gun, Model 981-2043 is installed for ion sputtering to get rid of impurities on the surface. A Dycor quadrupole 59 mass spectrometer, Model MA100M, is controlled through R5232 interface to measure Temperature programmed desorption (TPD) measurements and residual gas analysis. On the infrared level, there are two differentially pumped ports for CsI windows and two leak valves connected to the gas doser [19,20]. The spectrometer and the cryostat were home built at MSU. The system consists of a spectrometer shell, spectrometer top cover, liquid nitrogen can, spectrometer base plate and radiation shield. All the optical components, such as diffraction grating, folding mirrors, slit and filter wheels, the polarizer, and paraboloidal mirrors, are mounted on the base plate. The pressure of the spectrometer can be pumped down to 5x10"7 torr. The cryostat design is similar to the source cryostat. The grating drive mechanism is a combination of a stepping motor, a controller with a 360:1 worm gear, a 100:1 speed reducer. The speed reducer is engaged or disengaged to the driving shaft for fine or coarse grating movements using a computer controlled clutch. Precise angular position reading of the grating is possible through an lnductosyn rotary position transducer. A complete description is given in Ref. [6] The infrared beam, which is modulated by the tuning fork chopper, is collected by the detector and converted to an electrical voltage signal by the transirnpedence amplifier in the detector. The detector signal is demodulated by a PAR 5209 lock-in amplifier and read through a GPIB interface in an IBM 60 AT compatible computer. The computer can control the grating drive by communicating with the stepping motor controller and lnductosyn to read and change the grating angle with feedback and also permit precise data manipulation. The interfacing control program was written in Asyst programming language by Prof. R. G. Tobin. 2.7 Summary and discussion This chapter presents general ideas of a generic RAIRS systems. Two specific systems were described. Basic considerations are presented under idealized situations. Table 2.1 lists the practical numbers for these two systems. I have used these two systems to study adsorbate-induced broadband reflectance changes for atomic oxygen and formate on Cu(100) and CO on Pt(111). The measured signals are also listed in Table 2.1. Although the U4IR beamline can access longer wavelength and slightly better sensitivity, the MSU system is far more stable for broadband measurements. The numbers given are from the average performing results. The instability from source fluctuation and cryostat thermal drift are still expected sometimes during measurments. Global orbit feedback system has been installed in the storage ring at NSLS to stabilize the source fluctuation recently. And wavelength modulation is chosen for MSU system to promote the overall performance of the system. Further improvements in the instruments will bring more abundant results. 61 Table 2.1 The performance numbers of two RAIRS MSU. systems at NSLS and RAIRS System U4IR NSLS MSU Frequency range 80cm'1 to mid-IR 350 cm'1 to 2500 cm'1 Sensitivity 0.02% 0.02% rms Stability 0.2% over 20 mins 0.02% over 20 mins ReSOIUtion 1 CHI-1* 1 CHI-1 Adsorption Systems CO/ Cu(100) CO/Pt(111) O/Cu(100) Signal measured ~1.1% <0.3% * measured with a CsI window Refer s» 62 References 10. 11. 12. 13. 14. 15. GP. Williams, Nuclear Instruments and Methods, 195, 383 (1982); W. D. Duncan and G. P. Williams, Appl. Opt. 22, 2914 (1983). GP. Williams, Int. J. Infrared Millimeter Waves 5, 829(1984). G.P. Williams, Conceptual Design Report BNL #35985. C. J. Hirschmugl, G. P. Williams, F. M. Hoffmann and Y. I. Chabal, Rev. Sci. Instrum. 60, 2176 (1989). CJ. Hirschmugl, doctor’s dissertation, Yale University, 1994. C. Chung, doctor’s dissertation, MSU,1993. P.L. Richards and R.G. Tobin, in Vibrational Spectroscopy of Molecules on Surfaces, edited by ].T. Yates Jr. and TE Madey, pp. 413-483 (Plenum, New York, 1987) Y. J. Chabal, Surf. Sci. Reports 8, 211 (1988). M. A. Ordal, R. J. Bell, R. W. A. In, L. L. Long and M. R. Querry, Applied Optics 24, 4493 (1985). R. G. Greenler, J. Vac. Sci. Technol. 12, 1410 (1975). K. C. Lin, R. G. Tobin and P. Dumas, Phys. Rev. B 49, 7273 (1994). Laboratory Notes in 1993 of Dr. R. G. Tobin’s Lab. K.-]. Kim, Lawrence Berkeley Laboratory Report (1986). RJ. Bell, Introductory Fourier Transform Spectroscopy, Academic, New York (1972). A. Zangwill, Physics at Surface 5 (Cambridge University Press, 1988) 16. 17. 18. 19. 20. 63 Product and Vacuum Technology Reference Book, Leybold Vacuum Products, Inc. 1994 Perkin Elmer Physical Electronics, Auger Handbook. P.A. Redhead, Vacuum 12, 203 (1962). R. G. Tobin, C. Chung and]. S. Luo, J. Vac. Sci. Technol. A 12, 264 (1994). D. E. Kuhl and R. G. Tobin, Rev. Sci. Inst. 66, 3016 (1995). betwe resisti and C: comps; systerr chapte Chapter 3 Adsorbate-Induced Changes in the Infrared Reflectance and Resistivity of Metals In this chapter, I present an experimental test of the relationship between adsorbate-induced changes in the infrared reflectance and dc resistivity of a metal. The changes in reflectance for O/Cu(100), CO/Cu(100) and CO/Ni(100) over the frequency range 300-2000 cm’1 were measured and compared to the published resistivity changes for the corresponding thin film systems. These results were previously published in Ref. [1], on which this chapter is based. It has been widely known that adsorbates increases the dc resistivity p of thin metal films [2-4] and there have been several earlier attempts [5-7] to relate both optical (in visible light region) and resistivity measurements to scattering of conducting electrons, which modifies the dielectric properties of the substrate. The adlayer-metal interaction is taken into account through the p-parameter and the modified dielectric constants. The scattering is found to be one dominant factor in the experimental observation. We derive a simple formula to present the relationship based on the ideas of Persson’s model [8]. We intend to test the scattering hypothesis, by comparing our IR reflectance measurements with published data on thin film resistivity changes with this ancillary model. We emphasize that the 64 relatic- refererl conne. resona arein; but a I; 3.1 infrare Nation; to the: to freq- bombar combin to 11001 Cu 920- Was det 300Kfo' 65 relationship between resistivity and reflectance is very general; it makes no reference to the damping of the adsorbate’s hindered translation and the only connection to the proposed model of atomic friction and antiabsorption resonance is that they all invoke electron scattering. The experimental results are in good agreement to theoretic predictions for CO/Cu(100) and O/Cu(100), but a large discrepancy is found for CO/Ni(100). 3.1 Experimental technique The reflectance experiments were performed using the U4IR far- infrared beamline at National Synchrotron Light Source at Brookhaven National Laboratory. The light was p-polarized and incident at an angle of 85° to the surface normal. We used a Ge:Cu photoconductive detector sensitive to frequencies above 300 cm'l. The Cu crystal was cleaned using N e-ion bombardment at 500K followed by annealing to 750K, and the Ni crystal by a combination of ion bombardment at 600K, oxygen treatment, and annealing to 1100K. No surface impurity showed an Auger peak larger than 0.3% of the Cu 920-eV peak or 1.0% of the Ni 100-eV peak. For O on Cu(100), the coverage was determined from the oxygen Auger peak height, assuming a saturation coverage of 0.50ML at room temperature. For CO/Cu and Cu/ Ni, the coverage was determined by TPD and LEED. The sample temperatures were 300 K for CO/Ni(100) and O/Cu(100), and 90K for CO/Cu(100). 66 Fig. 3.1 (a) shows the fractional reflectance change ARp/Rp for two concessive spectra of a clean surface; the deviation from zero indicates the level of systematic error. Taking the uncertainty of the current and beam drift in the storage ring into account, we can determine ARp/Rp to an absolute accuracy of 1:0.2%. The spectra are composed from the measurements of two detectors. Spectra (b)-(d) of Fig. 3.1 show ARp/Rp for CO/Cu(100), O/Cu(100) and CO/Ni(100) over the frequency range from 300 to 2000 cm’l. The low frequency region in (d) is measured by a Si bolometer. The spectral resolution was 6 cm'1 for O/ Cu and CO/Ni and 16 cm'1 for CO/Cu. The data above 750 cm'1 have been smoothed to a resolution of 23 cm'l. The noise at the high frequency region and the absence of data in certain regions in spectra (a), (c) and (d) are due to the low transmittance of the polyethylene windows. Cesium iodide windows were used for spectrum (b). The data for CO/ Ni are almost consistent with zero reflectance change. The reflectance change seems slightly less than zero but shows no frequency dependence. The magnitude of the change is within the experimental uncertainty and much smaller than the observed values for CO/ Cu and O/ Cu. For 0.4 ML CO on Cu(100), ARp/Rp approaches zero at low frequencies and reaches a constant value of -1.1i0.2% above 1500 cm'l. This is consistent with a value of -1.9:t:0.1°/o measured at an incident angle of 87° by Borguet and Dai [9]. For 0.25ML O on Flg’Ur CO/r 101-m - 67 0.5 - 5 3 I 1‘ l.’ -; ; .." 00 _...- - M‘_IWME\II§- - -\,~‘;:“?‘:‘*3§1..:i; (a) clean '. ~05 * - 1 1 i 0.5 . . . 1 0.0 _. -------------------- (b) CO/Ni -0.5 * e a 0.5 - . 0.0 c..-.- - - --- _,-_..._, — -.- - -...... - .. . .0... \ . -1.0 - W WA: ARp/Rp(%) .0.— mu 1 .1 00 _.-.-.-....-.-.-.-.......-......._....-.-.-.-.l 0.5 e . .1 g > 1 = E ‘I ': " l S (d) O/Cu . " .- ' o 400 800 1200 1600 2000 Frequency ( crn'l ) Figure 3.1 ARp/Rp for clean Cu surface, CO/Cu(100), O/Cu(100) and CO/Ni(100) over the frequency range from 300 to 2000 cm'l. (modified form Fig. 1 in Ref. [1] ) (in 3511 exp can: 32 corr treq‘ isst to a para llere electr rehac thESL 68 Cu(100), the reflectance change shows the same characteristic as CO/ Cu. The asymptotic value is also about -1.1.t0.2% at high frequency. These reflectance changes are too large and of the wrong sign to be explained by the dielectric properties of the adsorbates [10]. They must be caused by the changes of dielectric properties of the metal. 3.2 Surface effect on reflectance The calculation of surface effects on reflectance requires nonlocal corrections to the classical Fresnel formulas [11,12]. We mainly focus on the frequency region where the local description of the metal’s dielectric response is still valid. The fractional change in the reflectance of p-polarized light due to a change of surface properties can be calculated using Feibelman’s d- parameter formalism [12]: ARL=_4(” Im(d,,) RP c c056 ' Eq. 3.1 provided that ls l >>(1/cosG)>>1, with e the complex (local) dielectric function of the metal, 0 the angle of incident, a) the frequency of the light, and d. I = [2(80. , / az)dz/] (80'. I / 3Z)dz Eq. 3.2 Here 0.. is the complex conductivity parallel to the surface for a uniform electric field, and z is the direction normal to the surface. (Because of strong refraction by the metal, the electric field inside the metal is nearly parallel to the surface, even for p-polarized light at near grazing angle incidence [13].) We 2155 Near t} allECl t ' tree pa Per C) v (I) (I) value conduc. the mc The co adsor": reflect: reSpoy ElEClTic QIECII'C llEQUe: Drude 69 We assume [8] that Im(d 1) is negligible compared to Im(d, I)- The problem of finding the reflectance change therefore reduces to that of calculating o. .(z). Deep inside the metal, the conductivity simply has its bulk value 03. Near the surface, it is modified by the presence of adsorbates. These changes affect the electron current to a depth on the order of the electron elastic mean free path lgzvprg (1);: is the Fermi velocity and the TB the bulk scattering time). Persson proposes a slab model [8] in which the conductivity has a constant value 05(m)=03(0))+AO'((D) from the surface to a depth 13, and the bulk conductivity on throughout the rest of the solid. When d I . is calculated from the model, it follows that: &=_aea_1_hn[em], Eq.... Rp c cos0 03(0)) The conductivity of a film of thickness 13 in this model is simply as, so if the adsorbate-induced change in conductivity of such a film is known, the reflectance change for the same adsorption on a bulk crystal can be calculated. It is appropriate to use a local description of the metal’s dielectric response, provided that 5>>uF/co, where 8 is the skin depth [8]; that is the electric field should not vary appreciably over the distance traveled by an electron during one cycle of the electric field. For both Cu and Ni in the frequency range studied, the local response is well described [14, 15] by the Drude model, in which the bulk conductivity is given by : where bulk r princip in the respect At zerc thick they 70 nezrB m(1—im3) ' 03(0)) = Eq. 3.4 where n is the conduction-electron density and m is the electron mass. The bulk material parameters are listed in Table 3.1. The adsorbate can in principle affect both n and rnear the surface. We can expand A6 to first order in the changes An and A1 in the electron density and scattering time, respectively: Ao(ro)_§£+ Ar 03(0)) n3 tau-ions) Eq. 3.5 At zero frequency, both terms enter equally, but only the term in Ar has an imaginary part that can give rise to a reflectance change through Eq. 3.3 (A second-order term in AnAt has been used to model absorption by a surface state [16,17]) More detailed treatments lead to essentially the same conclusion [5,7,15]. Ifwe assume that scattering dominates (An/n3<>1/1.'B, we find that : 1m[éa_(w_)] = _ 1 [AP (13 ):| Eq. 3.6 as ((9) 601' 8 p3 In fact, we can use the DC resistivity change for a film of arbitrary thickness t, since it is well established that Ap(t)oc1/ t [2,4]. The reflectance change can therefore be written as: pretax limit micro PIOPC indep optica contri’: scatter that d. can Wr This is and 49 dampii Ea cot the Sca Chapted 71 fl = 413 APUB) ___ ___4___ Rp CTB c050 p8 chrB c036 [tAp(t)]. Eq. 3.7 Only the term tAp refers to the thin-film sample; all of the quantities in the prefactor relate to the bulk crystal. The refelctance change at high frequency limit is predicted to be independent of frequency and proportional to tAp. A microscopic scattering calculation by Watanabe and Hiratuka also found ARP proportional to Ap [5]. For simplicity, we have derived this result asuming a frequency- independent scattering time 1'3, but the same equation is obtained if separate optical (top) and dc (tdc) [15] times are used, provided that the adsorbate contribution to the scattering rate is frequency independent. Then the DC scattering timet'dc appears in the denominator. But p31'dc=m/ne2 is a constant that depends only on the density of conduction electrons in the metal, so we can write Eq. 3.7: AR, _4ne2 1 [tAp(t)], Eq. 3.8 Rp mc c059 This is similar to Eq. 40 of Ref. [8]. It can also be obtained by combining Eqs. 41 and 49 of Ref. [8], but the present derivation avoids any assumption about the damping of adsorbate vibrations. We should note that the resistivity change is a coverage-dependent term. How the population of the adsorbates affects the scattering is an interesting question, and we will address it in the next chapter. 33 refle COUI Inea equi best Inaxi hon} avafle depei varie: 72 3.3 Comaprison between reflectance and resistivity We attempt to test Eq. 3.8 by comparing our measurements of IR reflectance with published data on thin film resistivity changes. It would of course be preferable to make simultaneous reflectance and resistivity measurements on well-characterized films deposited in situ, but we were not equipped to attempt this difficult measurement. In Table 3.2, we present the best available literature values for the resistivity change tAp, our measured maximum reflectance change ARP/ RP values, and reflectance change predicted from Eq. 3.8 using the electron density n values listed in Table 3.1. For 0 on Cu and CO on Ni, explicit plots of tAp versus coverage 11,, are available in the literature [3,18,19]. For CO on Cu, the shape of the coverage dependence Ap(na) is quite reproducible [2,6,7], while tAp at saturation value varies considerably. Wissmann [20] reports an average saturation value of tAp (for ~50 measurements) of 1.5x10'12 Qcm2 for films annealed at 60 °C, while a series of films annealed at room temperature gave values of tAp ranging from 1.1 to 2.6x10'12 (2cm2 [6]. The effect of annealing temperatures higher that 60 °C is unknown, as films develop cracks. To arrive at the values of tAp listed in Table 3.2, we have scaled the coverage-dependence curves by Wissmann’s average saturation value. lat elec con deg elet div J l j l Table 3.1 Bulk material parameters for Cu and Ni. Values for conduction electron density n and Fermi velocity 1),: are calculated assuming one conducting electron per atom [23,24]. The optical scattering time IB and skin depth are taken from Drude parameters given by Ordal et a1. [14]; the optical electron mean free path is given by IB=UPTB. Frequencies in s’1 have been divided by 27m to convert to cm'l. Metal n 73-1 DF/s 1., 5 (1022 CIn’3) (cm'l) (cm'l) (A) (A) Cu 8.5 73 310 1140 270 Ni 9.2 350 210 240 400 lat“. light I calci. thick 9:85 74 Table 3.2 Adsorbate-induced changes ARp/Rp in the reflectance of p-polarized light at the adsorbate coverage na as indicated, measured experimentally and calculated from Eq. 3.8 useing the indicated values of tAp for thin films of thickness t, the electron density n given in table 3.1, and an angle of incidence 6=85°. See the text for discussion of the uncertainties in tAp. System n, tAp Angry)“l A(RP/Rp)exp (10“ cm'zl (10‘12 (2 cm2) PM PM 7267511000) 6.1 0.603 -2.2 -1.1¢0.2 4.1101" —C-)/Cu(100) 3.8 0.42c -1.5 -1.1:tO.2 0.70“1 -2.6 "To/Niuom 6.4 1.47" -5.8 0210.2 3 See text b Reference [9], corrected for the difference in incident angles. ° Reference [3]. d Reference [18]. e Reference [19]. exper mode small mode greate mode for su for C a‘PPro 0» Tzl At 1m COUEC. 75 Variable sample preparation and uncertainties about thickness and surface sturcture make all the values of tAp reliable only within about a factor of two, as listed values for O on Cu suggest. Uncertainty also arises from the crystallographic orientation of surfaces; the films are polycrystalline, but are believed to be predominantly of [111] orientation [2], while (100) crystals were used for the reflectance measurements. This difference in orientation could affect the strength of electron scattering by the adsorbates. For the two adsorbates on Cu, both tAp and ARp/Rp are the same within experimental error. These results must be regarded as consistent with the model, even though the measured reflectance change is roughly a factor 2 smaller than predicted. This discrepancy could be due to the simplicity of the model, the uncertainty in tAp or the differences between the surfaces For CD on Ni, however, the predicted ARp/Rp is nearly a factor 30 greater than the observed value. Uncertainty in n, inadequacy of the slab model, and deviation from the Drude approximation are not likely to account for such a large discrepancy -- particularly since the agreement is much better for Cu. The disagreements cannot be assigned to a failure of the approximations leading to Eq. 3.8, even though the derivation requires m>>rg1 and vp/S, conditions that are only marginally met in our experiment. At 1000 cm'l, (ofi/vp~3 for Cu, and m3~3 for Ni (see table 3.1), the fractional correction to Eq. 3.8 due to these effects, however, is at most only ~10°/o; these 76 errors are much smaller than the uncertainty in tAp, and are roughly the same for Cu and Ni. For Ni, unlike Cu, lB<5; according to Ref. [11], this would render the d-parameter formalism invalid. But the formalism in fact requires only, IlEl<<8, where I: =(w+i/TB)/vp. For Ni above 1000 cm'l, lE~1)F/(D, so this requirement simply reduces to the locality condition a1>>vF/5 discussed above. According to the latest paper of the theory [21], the discrepancy for CO on Ni is explained by the scaling factor 5/1 which affects the characteristic of the universal function. Based on the free electron model and under the assumption of the same specularity caused by diffuse scattering from the adsorbates, the reflectance change for Ni would be a factor two small than for Cu. From Table 3.2, the reflectance change for CO on Ni is much less than 0.5%. The scaling effect is still too small to account for the discrepancy. Another possible explanation is that the reported tAp for CO on Ni is overestimated. Somehow the film may have been of nonuniform thickness; there are apparently no more recent experiments to compare with. However, the discrepancy between theory and experiment is so large that it suggests that for this system either scattering is not the dominant cause of the change in the resistivity or the theory relating reflectance and resistivity is seriously incomplete. Changes in n have been invoked to explain thin film resistivity changes due to surfaces, defects, and adsorbates, and oxygen incorporated at grain boundaries in thin Cu films has been shown [21] to reduce n ; such a mec Hm this 3.4 77 mechanism could in principle account for the small reflectance change. However, more experimental evidence is required to resolve the origin of this discrepancy. 3.4 Summary I have reported measurements of changes in the broadband IR reflectance for CO and atomic oxygen on Cu(100) and CO on Ni(100) and compared them with the changes predicted from published thin-film resistivity measurements, using a simple model based on the scattering of conduction electron by the adsorbates. For the adsorbates on Cu, the results are consistent with the model, but a large and unexplained discrepancy was found for CO on Ni. Uncertainties in the resistivity data , and doubts about the comparability of the thin-film and single crystal surfaces, however, prevent a definitive and quantitative test. There is a clear need for more experimental work. Reflectance and resistivity measurements should be performed simultaneously on well-characterized samples, such as epitaxial films. Refe 7m 78 References 1. 10. 11. 12. 13. KC. Lin, R.G.Tobin, P. Dumas, CJ. Hirschmugl, GP. Williams, Phys. Rev. B 48, 2791 (1992). P. Wissmann, in Surface Physics, edited by G. Hohler, Springer Tracts in Modern Physics Vol. 77 (Springer, New York, 1975). D. Dayal, H.-U. Finzel, and P. Wissmann in Thin Solid Films and Gas Chemisorption, edited by P.Wissmann (Elsevier, Amsterdam, 1987). D. Schumacher, Surface Scattering Experiments with Conducting Electrons, Springer Tracts in Modern Physics Vol. 128 (Springer, New York, 1975). M. Watanabe and A. Hiratuka, Surf. Sci. 86, 398(1979). U. Merkt and P. Wissmann, Z. Phys. Chem. Neue Folge 135, 227 (1983). M. Watanabe and P. Wissmann, Surf. Sci. 138, 95(1984). B.N.]. Persson, Phys. Rev. B 44, 3277 (1991); B.N.I. Persson and A. I. Volokitin, Chem. Phys. Lett., 185, 292 (1991). E. Borguet, I. Dvorak and H.-L. Dai, Laser Techniques for Surface Science 2125, 12 (1994). R.G. Tobin, Phys. Rev. B 45, 12 110 (1992). W.L. Schaich and W. Chen, Phys. Rev. B 39, 10714 (1989), and references therein. PJ. Feibelman, Prog. Surf. Sci. 12, 287 (1982). Y. I. Chabal, Surf. Sci. Reports 8, 211 (1988). 13. 17. 18. 19. 14. 15. 16. 17. 18. 19. 20. 21. 23. 24. 79 M. A. Ordal, R. J. Bell, R.W.A. In, L. L. Long and M. R. Querry, Applied Optics 24, 4493 (1985). F. Abeles in Optical Properties of Solids, edited by F. Abeles (John Wiley and Sons, New York, 1966). ].E. Reutt, YJ. Chabal, and SB. Christrnan, Phys. Rev. B 38, 3112 (1988). D.M. Riffe, L.M. Hanssen, and AJ. Sievers, Phys. Rev. B 34, 692 (1986); D.M. Riffe, L.M. Hanssen, and AJ. Sievers, Surf. Sci. 176, 679 (1986); D.M. Riffe and AJ. Sievers, Surf. Sci. 210, L215(1989) H. Buck, R. Schmidt, and P. Wissmann in Nichtrnetalle in Metallen, edited by D. Hirschfeld (DGM_Informationsgesellschaft, Wiesbaden, 1988), p.219. G. Wedler, H. Reichenberger, and G. Wenzel, Z. Naturforsch. 26, 1452 (1971). P. Wissmann, Thin Solid Films 13, 189 (1972); P. Wissmann (private communication). B. N. I. Persson and A. I. Volokitin, Surf. Sci. 310, 314 (1994). J. Vancea and H. Hoffmann, Thin Solid Films 92, 219 (1982). N. W. Ashcroft and N. D. Mermin, Solid State Physics (Saunders College, Holt, Rinehart and Winston, Philadephia, 1976). W. A. Reed and E. Fawcett, J. Appl. Phys. 35, 754 (1964). broac ofat. at to the I 0.25 l defin to be At 0. predi devia Surfac based 4-1 E) the \ Chapter 4 Adsorbate-Induced Changes in the Broadband Reflectance of a Metal: Oxygen on Cu(100) In this chapter I will present an investigation of the adsorbate-induced broadband reflectance change for atomic oxygen on Cu(100). Four coverages of atomic oxygen were measured through the frequency range 100 to 2000 cm'1 at room temperature. The observed frequency-dependence provides a test of the conduction electron scattering model which incorporates nonlocal electrodynamics. We find good agreement with theory for coverages below 0.25 ML. The value of the characteristic roll-off frequency (opvp/c, which is defined through conduction electron properties of the substrate and predicted to be independent of the adsorbate, is consistent for CO and oxygen on copper. At 0.35 ML coverage, the frequency dependence deviates strongly from the predicted form and the magnitude of the reflectance change decreases. These deviations may be related to an incipient ordering and reconstruction of the surface. These results have been published as Ref. [1], on which this chapter is based. 4.1 Experimental Procedures The experiments were performed at the U4IR far-infrared beamline at the National Synchrotron Light Source at Brookhaven National Laboratory. The light was p-polarized and incident at an angle of 85° to the surface 80 81 normal; these conditions are nearly optimal for the observation of reflectance changes due either to adsorbate vibrations or to changes in surface electronic properties [2]. Ultrahigh-vacuum-compatible polyethylene windows allow the spectral range to expand down to 50 cm'1 but have low transmittance above 750 cm'l. We measured the spectra from 100 to 2000 cm"1 using two detectors: a Si:B bolometer for frequencies below 650 cm'1 and a CuzGe photoconductor for frequencies above 300 cm'l. The nominal resolution was 6 cm'1 for the bolometer data and 16 cm'1 for the photoconductor data. During each run the sample temperature was stabilized to avoid thermally induced signal drift (a change in the temperature of 5 K produced observable intensity changes). The data presented here represent runs at varied sample temperatures between 300 and 330 K; the magnitude of the reflectance changes were found to be insensitive to the temperature within this range. The Cu(100) crystal was oriented within 0.5° and mechanically polished. It was cleaned by 500 eV N e+ ion bombardment at 600 grazing incidence to increase the efficiency and cover the whole surface more uniformly. The temperature was held at 550K for annealing during sputtering and annealed to a temperature between 700K and 850 K. Cycles of sputtering and annealing were applied until the contamination level was less than 1% of a monolayer. AES was used to monitor the contamination level. S, O and C are the most common impurities found on copper surface. No COVEI lh‘ca Tiiac Sh0w Obsex Overt annea We di 1'51“ 82 impurity revealed an Auger peak intensity larger than 0.3% of the Cu 920-eV peak (LMM). Dosing was accomplished by backfilling the chamber with Oz gas. At room temperatures 02 dissociates on the surface leaving an overlayer of atomic oxygen [3-10]. Heating and cooling the sample invariably resulted in severe baseline drift, so it was not annealed after dosing. The coverage of atomic oxygen on Cu(100) was determined from AES measurements. The Auger intensity ratio of the O KLL line (512 eV) to the Cu LMM line (920 eV) is proportional to the oxygen coverage. The characteristic oxygen coverage vs dosage curve is shown in Fig. 4.1, which presents our data for the ratio of Auger intensities of O/Cu as a function of dosage at 300 K. The saturation coverage for atomic oxygen is about one half monolayer at room temperature by calibrating the AES signal to a well known adsorption system, oxygen on Ni according to Wuttig, Franchy and Ibach [5]. For comparison, Fig. 4.1 also shows their results of the corresponding coverage vs. exposure data. We also conducted qualitative low-energy electron diffraction (LEED) observations. At the coverages and temperatures discussed here, no overlayer ordering was observed, which is consistent with other reports that annealing is required for optimum ordering. At higher coverages (> 0.45ML) we did observe a c(2x2) pattern, which according to Wuttig, Franchy and Ibach is in fact characteristic of the reconstructed phase. Flgm fimct obtai are Q COVe 0fth 83 0.20 v r I T I 5?: O ‘0 ‘ o o o 1 05 o .. O I ‘ v 0K4 ‘. A 5 015 ~ 2 h (:3 .-‘ . « 0.4 S 5 '0“ g 3 ‘ 009 7 Y -. E .3 010 P ' -'.3A S0.24 - 0.3 :3 1'! i '6 t 8 m 0'06’ "'5 «016 a b > . I3 . .1 02 0 t 003.” ' i” 3 0.05 ~. ‘ 39 +0.03] 5 S i l 01 L60 >3 0.000 15 30 ‘ 45 0°00 ' g S 0.0 0.00 A l A l l l i l A I 0 400 800 1200 1600 2000 Oxygen Dosage (L) Figure 4.1 Ratio of the oxygen KLL to copper LMM Auger intensities as a function of oxygen dosage at 300 K. The coverage scale on the right axis is obtained by assuming a saturation coverage of 0.50 ML. The solid symbols are our measurements on two separate days; the open circles show coverage vs. exposure data from Ref. [5]. The inset shows an enlargement of the low-dosage region. (from Ref. [1]) 4.2 Data Reduction The intensity of synchrotron radiation is proportional to the electron current in the storage ring [11]. All the spectra were normalized to the beam current before analysis. The normalization process is: at the beginning of each scan, the time was recorded by the data acquisition system. The beam currents were also recorded periodically. The normalizations were performed after the experiments were completed. The beam current was derived for the starting time for each spectrum by interpolation from a least-squares linear fit to the measured currents. Although the beam decays exponentially due to loss of electrons in the stored beam, a testing result of this normalization shows that the reproducibility was better than 0.2% over a 20 min. period with a CuzGe detector. For a Si bolometer detector, the reproducibility is slightly worse which results from the strong dependence of its responsivity to temperature. The overall systematic error of the absolute reflectance measurements is about 0.2%. The uncertainty of the beam current and the drift of the stored electron beam position, however, can also cause intensity variations that mimic reflectance changes. To exclude such spurious effects, we adopted a stringent and unbiased system of data selection and analysis, as described in the following. 85 After the sample had been cleaned and its temperature was stable, four to eight consecutive reference spectra were measured. Each spectrum comprised 128 scans and took 55 sec. The sample was closed with oxygen and another set of spectra was measured. Ratios of the individual, current- normalized reference spectra with each other were examined, as well as ratios among the individual post-dosing spectra. If any of these ratios of nominally identical spectra showed baseline deviations greater than 0.2% between 300 and 2000 cm'l, the entire experiment was discarded. Comparisons of the post- dosing spectra with the reference spectra were not used in selecting the data. Approximately one-third of the measurements met the stability criterion, leaving three photoconductor and two bolometer runs for 0=0.25ML and one run with each detector at 0.08, 0.17 and 0.35 ML coverages. Because of the long dosing time required, spectra at higher coverages invariably showed excessive drift. The reference spectra (R) and post-dosing spectra (R’) from each run were averaged. Then the fractional reflectance change AR/R (AR=R’-R) was calculated. Owing to systematic errors, there was in general an offset (typically ~0.05°/o) between the bolometer and photoconductor spectra in the frequency range (400-650 cm'l) covered by both detectors. To obtain a final composite spectrum the bolometer data were offset to agree with the photoconductor signal. For 0 =0.08 and 0.35 ML, baseline deviations were found below 300 cm' 1 and the low-frequency region was discarded. 86 As a final check against instrumental drift, the oxygen dosage was increased to 50L and the sample was exposed to 50L of formic acid, which removed the adsorbed oxygen, leaving a layer of formate species on the surface [12]. The reflectance change AR/R (relative to the clean surface) returned to zero within 0.2%. It clearly demonstrates that the reflectance changes shown are in fact oxygen induced. For oxygen coverages >0.25 ML formic acid does not completely remove the oxygen and this test could not be used. 4.3 Experimental results Fig. 4.3 presents the composite reflectance spectra for four coverages of O on Cu(100). The gaps in the spectra and the noise at high frequencies are due to the low transmittance of the polyethylene windows. All show the same qualitative behavior-- the reflectance change is negative, and its magnitude is small at low frequencies, increases rapidly between 200 and 600 cm'l, and reaches an asymptotic value above 1500 cm'l. Theoretical considerations require AR/R to approach zero at zero frequency, as the wavelength and the penetration depth of the field become infinite [13—16]. We therefore attribute the apparent nonzero intercepts to systematic errors. The solid lines are fittings to the scattering model and will be discussed in next section. 87 Figure 4.2 Oxygen-induced fractional reflectance change AR/ R for four oxygen coverages: (a) 0.08 ML; (b) 0.17 ML; (c) 0.25 ML; (d) 0.35 ML. The gaps in the spectra and the noise at high frequencies are due to the low transmittance of the polyethylene windows. The solid lines are fits to Eq. 4.2 discussed in section 4.4; for spectrum (d) no fit was possible. AR/R AR/R AR/R -0.005 -0.010 -0.005 -0.010 r -0.015 -0.005 -0.010 - -0015 ' -0.020 0.005 _ -0.010 - -0.015 0 0.000 l _ 0 : t :7 °‘, 4 (a) e ‘- 0.08 . B ‘ e .- .1 0.000 T 3 ‘3; '53! v‘ :P: ®)0=017 1 ;. o t 0.000 ¢)e=025 0.000 (d) 0 = 0.35 I T I I I o . I i O ' Eel t ‘. ‘ . 3 53 '5 0.. e .’ a ‘ 3 t‘. "-‘l' ' -r .00... U h.‘ t}. ’0. 0” k ‘ .1-r. -.° 0.. ‘ o : .0 8' : 800 1200 Frequency ( cm'I ) 400 1600 2000 COI‘ llllt the lllE‘t 4.4 m0< the and diff the w} the th. be 89 The Cu-O stretch vibrations are expected near 340 cm'1 at low coverage and near 290 cm"1 at higher coverage [4,5,13]. We did not observe a convincing Cu-O stretch signal despite a sensitivity of 0.02%. This information would place an upper limit to the dynamic dipole moment of the vibration. I will make a comparison with other experimental and theoretical results in section 4.6. 4.4 Frequency Dependence In a series of papers [14-17], Persson and Volokitin have elaborated a model of adsorbate-induced-reflectance changes, based onthe assumption that the electron transport in the near surface region is perturbed by foreign atoms and the optical response to the incoming IR field drops due to the increase of diffuse conduction electron scattering by disordered adsorbates. As I discussed in section 1.4, the universal function for adsorbate-induced reflectance changes, valid for free electron metals at [R frequencies w>>r'1, is: 2 A; —4a (co/a2.) Re[g(w/wpl/5] , Eq. 4.1 — 732 40 ~ dq 1—[7rcop cosellm-L e(w/w,,l/6,q) where g(x,z) and e(x,z,q) are complex-valued functions defined in Ref. [17], l is the electron mean free path, 0),, is the plasma frequency of the metal, 6=c/(0p is the classical skin depth, and (01:60pvp/C is a characteristic frequency, near and below which nonlocal effects become important. wh Sin Nc bu inc C111 90 One of the theoretical predictions tested here is that only the magnitude a should depend upon the adsorbate or the coverage; the spectral shape and the parameter (01 should be properties only of the substrate. Applying Eq. 4.1 to the data of Hirschmugl et al. for CO on Cu(100), Persson found a good fit with a value of (01:500 cm'l, on the same order as theoretical estimates of 444 cm‘1 from a jellium model; an analysis of newer data found a value of 400 cm"1 [18] Equation 4.1 is computationally very cumbersome, and we did not attempt to fit it systematically to our data. Over the frequency range studied here, it is extremely well approximated by a much simpler expression, AR ~ 2 "R-=C-a‘—£6—, Eq.4.2 (02 +362 where for l/6=4.2, appropriate for copper at room temperature, the parameter sin Eq. 4.1 and 4.2 are related by: C = -—0.00064a , a = 0.96a , Eq.4.3 a = 1.21ch1 . Note that Eq. 4.2 is very similar to the approximation proposed by Persson building on the work of Dingle [19]. I have performed weighted least-squares fits of Eq. 4.2 to my data, including an additional offset C’ to account for systematic errors in the current normalization. The solid lines in Fig. 4.2(a)-(c) show the results and lahl Eq. - COVt exa: QUE COX“. see- the Son thec nea zerc all? Wh inn for Wit V0] 91 Table 4.1 gives the values and the uncertainties of the parameters a and (01 in Eq. 4.1 obtained through Eq. 4.3 and of the instrumental offset (C'-C), for each coverage at which a fit was possible. The uncertainties were estimated by examining the contours near the minimum x3 and for normally distributed errors represent better than 99% confidence limits [20]. The values of XE indicate an excellent fit for the two lowest coverages. The fit to the 0.25 ML coverage is still good, but there is a small but significant discrepancy in the 800-1200 cm" region. Fig. 4.3 compares Eq. 4.2 (solid line) with Eq. 4.1, using parameters related through Eq. 4.3 for our data for 0:0.25 ML with C’ subtracted. Within the accuracy of our measurements, the two curves are indistinguishable. Some disagreement is apparent at very low frequencies, where the general theory predicts that the reflectance change goes to zero with nonzero slope but near-zero curvature. The data, on the other hand, seem to be approaching zero slope with a distinct negative curvature--similar to the behavior of the approximate expression. A similar trend is visible in the data for CO [15,21]. While the discrepancies are too small to be conclusive, they may hint at an incompleteness in the theory at very low frequencies, and suggest the need for further theoretical work. For the three lowest coverages the excellent agreement of Eq. 4.1 and 4.2 with the data provides strong support for the model of Persson and Volokitin. 92 Table 4.1 Best-fit parameter values and uncertainties obtained by fitting Eq. 4.2 to the spectra in Fig. 4.2, and transforming the fitting parameters according to Eq. 4.3. The weighted average of the (01 values is 3384.34 cm'l. Also listed are the values of a and ml=mpvp/c found for CO on the same Cu(100) crystal, and the value of a); predicted from a free electron model. O/Cu(100) a(%) 601(cm'1) C’-C(%) 15 coverage 0.08 $47333 5553;: 4.15333 1.06 0.17 4.6333; 3:31:57; -0.163;3§ 1.10 0.25 4.04:3;3 339:3; 4.233% 1.34 Wuflm) 0.4“ -1.23 500 0.5b 4.33 400 Predicted 444c 310d aReference [21] and [16] bReference [18] cReference [15], based on one electron per Cu atom. dReference [24], based on an electron density deduced from optical measurements. 93 0.000 0004 g -0.008 <1 ., -0012 ~ ' " . f ';. 0020 L ~ 1 l ' 400 800 1200 1600 2000 Frequency (cm-l ) Figure 4.3 Oxygen-induced reflectance change for 0:20.25, with the instrumental offset C'-C subtracted. The solid line shows the best fit to Eq. 4.2 and the dashed line shows the results of Eq. 4.1 using the parameters given by Eq. 4.3. Over the frequency range studied, the two curves are virtually indistinguishable. 94 The values of (01 for the three coverages are also consistent. (The large uncertainty in an for 0:0.08 ML is caused by the weak variation of the reflectance and by the absence of data at low frequencies.) Assuming that (01 is independent of coverage- as predicted by the theory— the weighted average for the three curves gives (01:3382t34cm'1. It is surprising that this estimate differs from the values of 500 and 400 cm'1 found for CO on exactly the same Cu(100) crystal, since wlzwva/c should be independent of the adsorbate and the temperature. The high-coverage (0.35ML) data could not be fit by Eq. 4.2 (The minimum 13 value was 10.2). Since the form of the frequency dependence depends only on the assumption that the dominant effect of the adsorbate on conduction is increased scattering, this deviation suggests that other processes may be important at high coverages. Within the scattering model (see section 1.4, the magnitude a of the asymptotic reflectance change at high frequencies (a)>>a)1) is related to the change of the electron relaxation time which results in the change in dc resistivity Ap of a thin film of the same metal according to: _ 4ne2 l [tAp(t)], Eq. 4.4 mc c089 where n is the conduction electron density, 111 is the effective mass, t is the thickness of the metal film, and 6 is the angle of incidence. The quantity tAp 95 is independent of t for a given adsorption system [22,23]. In the previous chapter, we compared the reflectance change at high frequency (>1500 cm'l) for single coverages of oxygen and CO on Cu(100) and of CO on Ni with the values predicted from Eq. 4.4, using published data for the resistivity change of thin films [25]. Given the large variations and uncertainties in the resistivity data, and in the comparison between measurements on dissimilar samples, the agreement was reasonably good for the adsorbates on Cu. For CO on Ni(100) [24], however, the predicted reflectance change was large (~-6%) while essentially no reflectance change (<0.4%) was found experimentally. A qualitative explanation has been offered [16,17] that can account for roughly a factor of 2 difference between the predicted and measured values, which is similar to what is observed for CO on Pt and will be discussed in chapter 6, but the factor of 30 difference is too large to explain. Testing the scattering model by focusing on the frequency dependence of reflectance change is more convincing, since it avoids the large uncertainties in the resistivity data. In the next section, I also examine the dependence of the reflectance change on coverage. 4.5 Coverage dependence An important consequence of the scattering model is a qualitative prediction of the dependence of the reflectance or resistivity change on adsorbate coverage. At low coverage the adsorbates are randomly distributed, in the absence of long-range interactions, and the resistivity or reflectance 96 change should vary linearly with coverage. At higher coverage, the adsorbates may form an ordered overlayer. The effect of ordering on the resistivity and reflectance is an unsettled issue both theoretically and experimentally. No diffuse scattering is possible from an ordered surface, but Bragg scattering will contribute to the resistivity if the unit cell of the overlayer is larger than that of the clean surface. We would expect the Bragg scattering contribution to be less than that from diffuse scattering from the same adsorbate coverage, so that ordering should produce a decrease in the resistivity and in the magnitude of the reflectance change. Resistivity measurements on the films sometimes exhibit a maximum as a function of adsorbate coverage [25], but the resistance does not approach the clean-surface value and there is no direct evidence associating the decrease in resistance at high coverage with ordering. The roughness of the film is difficult to control and therefore the overlayer structure is not well defined. An IR reflectance study of H on W(100) by Riffe and Sievers [26] found a decrease in the magnitude of the reflectance change upon ordering, as expected. For CO on Cu(100), however, Hirschmugl et a1. observed a linear increase with no detectable change in slope as the ordered c(2x2) structure formed [27]. Figure 4.4 presents our data for the asymptotic (high-frequency) magnitude of the fractional reflectance change as a function of coverage. For the three lowest coverages, the values and uncertainties are those listed in 97 1.5 . . . < 0.16 33: 1.2 e 1 \ L O 3;“ 4* - 0.12 O 30 0.9 l 1 o 1 u "“ - 0.03 3 0.6 ~ 0 '21:." l C s + _23 o ‘15 0.3 ~ ‘ 0'04 05 0.0 I . 0.00 0.00 0.15 0.30 0.45 0.60 Oxygen Coverage (ML) Resistivity Change An (1 (Milan) Figure 4.4 Asymptotic (high frequency) limit of the reflectance change as a function of coverage (filled square). The values and uncertainties are taken from Table 4.1 for the three lowest coverages, and estimated from Fig. 4.2(d) for 020.35 ML. For comparison, the open circles show the resistivity change of a 40 nm-thick copper film (from Ref. [25] ). 98 Table 4.1 for the parameter a, which is the magnitude of the reflectance change. For 0:0.35 ML, we have made an estimate from the high-frequency portion of Fig. 4.2 (d). The uncertainty in the coverage includes both errors in dosing and uncertainty in the coverages vs. dosage curve in Fig.4.1. A qualitative change in behavior at the highest coverage is apparent both from Fig. 4.4 and from Fig. 4.2; the magnitude of the reflectance, rather than continuing to increase, drops about 20% from its value at 0.25 ML coverage. At the same time, the shape of the frequency dependence undergoes a marked change, reflected in the inability to fit the 0.35 ML coverage data, even approximately. We suggest that this change of behavior is related to the tendency of oxygen on Cu(100) to form an ordered, reconstructed surface at a coverage close to 0.35 ML. A number of experiments have established the existence of a (2J2 x «[2- )R45° structure involving a ”missing row” reconstruction [5-10]: every fourth row of Cu atoms is vacant. According to the LEED study of Wuttig, Franchy, and Ibach [5], there is a first-order phase transition from a disordered layer into the ordered phase at critical coverage 0c=0.34 ML, close to the highest coverage we were able to study. Our own LEED observation did not reveal the (242 x J2 )R 45° pattern at the coverage studied, probably because we were not able to anneal the overlayer (heating and cooling the sample invariably resulted in baseline 99 drift). Equilibration of the overlayer and formation of the ordered phase are known to require annealing to at least 400K. It is possible that some partial or local ordering was occurring at room temperature, which could not be observed with LEED but had a noticeable effect on the reflectance change. Qualitatively, we would expect some reduction in diffuse scattering for ordering on a length scale greater than the electrons, Fermi wavelength, approximately 5 A for copper; it is therefore plausible that ordering on a length scale too short to be observed with LEED could have an observable effect on the reflectance. We present this possibility as a speculation; clearly additional experiments are needed for a definitive conclusion. In Fig. 4.4, the coverage dependence of the resistivity change observed for oxygen on a thin copper film is also displayed. The three lowest coverage points behave similarly to the resistivity, with the reflectance change increasing monotonically, but with decreasing slope. The drop in the magnitude of the reflectance change at 0.35 ML coverage is not mirrored in the resistivity curve. The resistivity curve does exhibit a smaller (~10%) drop at higher coverage (> 0.6 ML, not shown in figure). If the drops are associated with surface ordering, it is not surprising that two curves are different since the surfaces are dissimilar (the films are heterogeneous, but expected to be dominantly of [111] orientation [2]). Alternatively, it is possible that the actual adsorbate coverages in the resistivity experiments were substantially lower than estimated. 100 4.6 Dynamic dipole moment of oxygen on Cu(100) We did not observe a convincing Cu-O stretch signal, despite a sensitivity of 0.02%. Figure 4.5 shows the spectreum for 0.25 ML oxygen on Cu(100) with the frequency range from 100 and 600 cm’1 after removing a baseline disscussed in previous sections. Using an estimated linewidth of 50 cm'l, we find the dynamic dipole moment e"<0.12 e. This compares with values from 0.12 e to 0.18 e estimated from electron energy loss spectroscopy intensities [28,29], assuming typical values of 1-1.5° for the spectrometer acceptance half-angle [29,30]. (Details of the calculation formalism are in Appendix A.) Possibly the lines were broadened in our experiment because the sample was not annealed, or part of the EELS intensity may result from impact scattering. The effective charge value for the Cu-O mode is rather small compared to other adsorbate-substrate vibrations [29]. This may be understood qualitatively in terms of screening of the incident electric field by the copper conducting electrons. Structural studies have found that the oxygen is less than 0.25 A above the center of the top Cu layer, a position which according to jellium calculation [31] places it well inside the screening electron density. However, a theoretical calculation of e" that include the effects of screening gives much larger values, in the vicinity of e*=1.0 8 [32,33]. This implies stronger polarization of O—Metal bonding. The large discrepancy between 101 0.0005 ‘ l ‘ T I I I l . u . v 0.0003 ~ , ; . " .. r' . . 1 0.0001 ~° 1...“... . . ; ’1'. ".;.;.' - g -.,' -.--j.j'..:; 15's: -0.0001 -; " if}? : xyfif 2. '24:. 5' 2° ° ’ l -000003 '- .".:: _. " 0 = 0.25 ML 0000? 00 ‘ 200 300 400 500 600 Frequency ( cm-l ) Figure 4.5 The low frequency spectrum from Fig. 4.2 (c) for 0.25 ML oxygen on Cu(100) surface after removing the electron-scattering induced background reflectance change. The O—Cu stretch is observed around 340 cm". 102 experimental results and theoretical predictions merits further theoretical investigation. 4.7 Conclusion Ihave measured the adsorbate-induced change in infrared reflectance over a wide frequency range (100-2000 cm'l) for oxygen on Cu(100) near room temperature. The data offer a test of a proposed model of surface scattering that incorporates nonlocal electrodynamics. For coverages below 0.25 ML, the data strongly support the model. A qualitative change in behavior is observed at the highest coverage studied (0.35 ML), where the magnitude of the reflectance change drops and the shape of the frequency dependence changes markedly. These changes may be associated with an incipient ordered phase, although annealing the surface to obtain long range ordering could not be obtained for experimental reasons. Further experimental and theoretical work is called for to clarify the effects of ordering, the generality of the scattering model, and the very-low-frequency behavior of the reflectance change. References 1. K. C. Lin, R. G. Tobin and P. Dumas, Phys. Rev. B 49, 17 273 (1994). 2. Y.]. Chabal, Surf. Sci. Rep. 8, 211, (1988) 3. A. Apitzer and H. Liith, Surf. Sci. 118, 121 (1981). 4. M.H. Mohamed and LL. Kesmodel, Surf. Sci. 185, L467 (1987) F‘ 5. M. Wuttig, R. Franchy, and H. Ibach, Surf. Sci. 213, 103 (1989). i 6. HO Zeng, R.A. McFarlane, and K.A.R. Mitchell, Surf. Sci. 208, L7 (1989). 7. I.K. Robinson, E. Vlieg, and S. Ferrer, Phys. Rev. B 42, 6954 (1990). in: 8. MC. Asensio, M.]. Ashwin, A.L.D. Kilcoyne, D.P. Woodruff, A.W. Robinson, Th. Linder, ].S. Somers, D.E. Ricken, and A.M. Brawshaw, Surf. Sci. 236, 1(1990). 9. Ch. W611, R. J. Wilson, S. Chiang, H.C. Zeng, and K.A.R. Mitchell, Phys. Rev. B 42, 11 926 (1990). 10. R. Mayer, C.-S. Zhang, and K.G. Lynn, Phys. Rev. B 33, 8899 (1986). 11. W.D. Duncan and G.P.Williams, Appl. Opt. 22, 2914 (1983). 12. I. Stohr, D.A. Outka, R.]. Madix, and U. Dfibler, Phys. Rev. Lett. 54, 1256 (1985). 13. BA. Sexton, Surf. Sci. 66, 56 (1977). 14. B.N.]. Persson, Phys. Rev. B 44, 3277 (1991). 15- B.N.]. Persson, Chem. Phys. Lett. 197, 7 (1992). 16. B.N.]. Persson and A. I. Volokitin, J. Electron Spectrosc. Relat. Phenom. 103 64/65, 23 (1993) 17. 18. 19. 20. 21. 23. 24. 26. 104 B. N. I. Persson and A. I. Volokitin, Surf. Sci. 310, 314 (1994). C. I. Hirschmugl, G. P. Williams, B. N. I. Persson and A. I. Volokitin, Surf. Sci. Lett. 317, L 1141 (1994). RB Dingle, Physica 19, 729 (1953). RR. Bevington and D.K. Robinson, Data Reduction and Error Analysis for the Physical Sciences, 2nd ed. (McGraw-Hill, New York, 1992), pp- 69-70, 258 and 259. C. I. Hirschmugl, G. P. Williams, F. M. Hoffmann and Y. I. Chabal, Phys. Rev. Lett. 65, 480 (1990). P. Wissmann, in Surface Physics, edited by G. Hohler, Springer Tracts in Modern Physics Vol. 77 (Springer, New York, 1975). D. Schumacher, Surface Scattering Experiments with Conducting Electrons, Springer Tracts in Modern Physics Vol. 128 (Springer, New York, 1975). K. C. Lin, R. G. Tobin, P. Dumas, C. I. Hirschmugl and G. P. Williams, Phys. Rev. B 48, 2791 (1992). D. Dayal, H.-U. Finzel, and P. Wissmann in Thin Solid Films and Gas Chemisorption, edited by P.Wissmann (Elsevier, Amsterdam, 1987); H. Buck, R. Schmidt, and P. Wissmann in Nichtmetalle in Metallen, edited by D. Hirschfeld (DGM_Informationsgesellschaft, Wiesbaden, 1988), p.21; G. Wedler, H. Reichenberger, and G. Wenzel, Z. Naturforsch. 26, 1452 (1971). D.M. Riffe and AJ. Sievers, Surf. Sci. 210, L215(1989). 27. 28. 29. 30. 31. 32. 33. 105 C. I. Hirschmugl, Y. J. Chabal, F. M. Hoffmann and G. P. Williams, J. Vac. Sci. Technol. A 12, 2229 (1994). S. Anderson and ].W. Davenport, Solid State Commun. 28, 677 (1978). H. Ibach, Surf. Sci. 66, 56 (1977); A.M. Barc’), H. Ibach, and H.D. Bruchman, Surf. Sci. 88, 384 (1979). M. Persson, S. Andersson, and PA. Karlsson, Chem. Phys. Lett. 111, 597 (1984). N.D. Lang and W. Kohn, Phys. Rev. B 7, 3541 (1973). PS. Bagus and F. Illas, Phys. Rev. B 42, 10 852 (1990). EA. Calhoun and ].E. Inglesfield, Phys. Rev. Lett. 66, 2006 (1991). Chapter 5 Infrared Spectroscopy of Formate on Oxygen-Predosed Cu(100): Broadband Reflectance and Low-Frequency Vibrations In this chapter I will discuss the adsorbate-induced broadband infrared reflectance change and low frequency vibrational modes of formate on an oxygen-predosed Cu(100) surface at room temperature. In contrast to CO, O, and H on the same copper crystal surface, no broadband reflectance change is detected over the frequency range 200-2000 cm'l. The difference may be associated with the lack of a formate-derived. electronic state near the Fermi level. Two vibrational bands are observed, at 365 and 380 cm‘l, attributed to oxygen-metal stretch vibrations; previous electron energy loss spectroscopy at lower resolution detected only a single band. We suggest that the doublet may indicate the coexistence of two formate species with different binding sites. These results have been published in Ref [1’], on which this chapter is based. 5.1 Introduction Interest in the chemisorption of formate (HCOO) on copper single crystals arises from its importance in heterogeneous catalysis: Formate is well-established as a surface intermediate in methanol synthesis and the catalytic decomposition of formic acid over copper surfaces. This system has been studied by a variety of surface sensitive techniques such as XPS, 106 107 NEXAFS, etc. and also investigated theoretically [2-17]. However, such basic issues as the binding site and adsorption geometry are still controversial. We present a reflection-absorption infrared spectroscopy (RAIRS) study of formate on oxygen-predosed Cu(100), using synchrotron radiation to span the frequency range from 200 to 2000 cm'l. There have been previous vibrational studies of formate on Cu(100) using electron energy loss spectroscopy (EELS), with lower resolution [1], and a RAIRS study of formate on Cu(110) by Hayden et al. [5], restricted to frequencies above 1300 cm‘l. To our knowledge this is the first reported RAIRS study of formate on Cu(100) and the first far-IR study of adsorbed formate on any surface. We find that the low-frequency metal-formate stretch vibration previously detected with EELS is actually a doublet, suggesting that formate may occupy two binding sites. We are also able to investigate the formate-induced change in the broadband IR reflectance of the surface. It has been shown that CO [18,19], oxygen [20,21], and hydrogen [2] all reduce the broadband reflectance by ~1% changes that are too large and have the wrong sign to be explained by the dielectric properties of the adsorbate layer [23]. Persson and Volokitin [24,25] have attributed the change, together with an antiabsorption feature observed for CO [18], the damping of adsorbate vibrations parallel to the surface and adsorbate-induced changes in the resistance of thin films, to diffuse scattering of conduction electrons by the adsorbates. Their formalism incorporates nonlocal electrodynamics to predict the frequency dependence of the 108 broadband reflectance and the line shape of the antiabsorption feature [25] and is in good agreement with experiment [19,20], at least for low coverages, where the adsorbates are randomly distributed. These comparisons between experimental results and theoretical predictions have been discussed in previous chapters. We find that formate, unlike these other adsorbates, produces a negligible change in the broadband reflectance. 5.2 Experiments The experiments were performed at the U4IR beamline at the National Synchrotron Light Source at Brookhaven National Laboratory. Details of the optics have been described elsewhere [26,27]. The light was p-polarized and incident at an angle of 85° to the surface normal. Polyethylene windows on the UHV chamber permitted the use of frequencies down to 50 cm'l, but had low transmittance above 750 cm'l, which resulted in noise and gaps in the spectrum. We used two detectors: a Si:B bolometer for frequencies below 550 cm'1 and a Ge:Cu photoconductor for frequencies above 300 cm‘l. The internal vibrational modes of formate, near 760 cm"1 and above 1000 cm‘1 lie in regions of low window transmittance. As a result, only the Cu-formate stretch mode was studied. The nominal resolution was 6 cm'1 for the bolometer data and 16 cm'1 for the photoconductor data. All spectra were normalized to the electron beam current in the synchrotron ring. During each run the sample temperature was stabilized to avoid thermally induced signal drift (a change in temperature of 5 K produced an observable intensity 109 change). The data presented here were taken near room temperature. A monolayer of formate is reported to be stable up to 400 K [2]. The sample was cleaned by Ne ion bombardment at 550 K, annealed to about 800 K, then cooled to the experimental temperature. No impurity showed an Auger signal larger than 0.3% of the Cu 920 eV peak. The formic acid gas was purified by repeated freeze-pump-thaw cycles. Although the formate species can also be obtained through direct decomposition of formic acid on Cu(100), predosing the surface with oxygen increases the sticking coefficient of formate. In these experiments the sample was first dosed with 50 L (1 L = 10‘6 torr sec) oxygen, giving an oxygen coverage of 0.25 ML, then exposed to 50 L of formic acid, giving a saturation coverage of formate. The absolute formate coverage was not determined. Temperature-programmed desorption (TPD) was used as an additional check on the formate layer formation. 5.3 Results and discussions 5.3.1 Broadband reflectance change. Great care is required in the analysis of broadband reflectance data to avoid systematic errors due to drift of the stored electron beam, uncertainty in the beam current, instrumental instabilities, and temperature fluctuations. Our procedures are fully described in chapter 4. Approximately one third of 110 the experimental runs met our unbiased stability criteria and are included in the analysis. Figure 5.1 compares the broadband reflectance change for 0.25 ML of oxygen with the spectrum obtained after subsequent exposure to 50 L of formic acid. The solid line through the oxygen spectrum is a fit to the scattering theory of Persson and Volokitin [25], as detailed in chapter 4. The formate spectrum shown in Fig. 5.1(b) combines the average of four separate photoconductor spectra with the average of four bolometer spectra; the averaged bolometer spectrum was offset by 0.07% to match the photoconductor spectrum between 340 and 400 cm'l. There is more systematic uncertainty in the formate data than in the oxygen or CO measurements because of the longer time required for the UHV system pressure to recover after dosing. Nevertheless it is very clear that replacing oxygen with formate dramatically reduces the magnitude of the broadband reflectance change. In fact the formate-induced reflectance change (relative to the clean surface) is smaller than 0.3%, and consistent with zero. Apparently formate is a much weaker scatterer of conduction electrons than oxygen, CO or hydrogen. It is initially surprising that formate should be such a weak scatterer. It is stable on Cu(100) to higher temperatures than either CO or H, indicating relatively strong bonding; theoretical estimates of the binding energy bear out this conclusion [14,17]. In addition, the formate moiety is expected to carry a 111 0.005 a E O. . .l‘ tfiht‘sw 0.000 ""' - --- "1M 3% _________ L-h!:1-:1 -0.005 AR/R -0.010 -0.015 . -. .3. -0.020 .1. 1214'!- 0 400 800 1200 1600 2000 Frequency( cm-l ) Figure 5.1 Adsorbate-induced fractional reflectance changes (AR/. R) due to (a) 0.25 ML oxygen on Cu(100); (b) formate on Cu(100) obtamed by exposing the oxygen-predosed surface to 50 L formic acrda The. data reduction procedures are discussed in the text. The SOlld line m (a) IS the best fit to the electron scattering model of Ref. [25]. The gaps m the spectra and the noise at high frequencies are due to the low transmittance of the polyethylene windows. 112 large negative charge [14], which might be expected to lead to strong scattering. A possible explanation is found in the electronic structure. Persson postulates [24] that scattering is strongly enhanced by the presence of an adsorbate- induced electronic state overlapping the Fermi level. For oxygen on Cu(111) and Cu(110), an inverse photoemission study by Jacob et al. [28] shows an adsorbate-induced state about 3 eV above the Fermi level which could extend into the Fermi energy region. Several experiments that probe the electronic structure of adsorbed CO indicate the presence of states within 2-3 eV of E; [29]. For formate on copper surfaces, however, the highest occupied state detected with ultraviolet photoelectron spectroscopy is about 5 eV below E; [11,12]. There appears to be neither experimental nor theoretical information regarding the unoccupied states. The absence of states near E; may accounts for formate's weak effect on the surface reflectance. 5.3.2 Copper-formate vibrations It is established that formate bonds to Cu in a bidentate configuration through its two oxygen atoms, but the exact binding site and geometry are still debated [3,4,7,8,9,14,16,17]. No ordered low energy electron diffraction pattern has been reported for this system, indicating that no long-range ordered structure is formed. It has been assumed that only one binding site is occupied, based in large part on EELS measurements showing single bands corresponding to the Cu-formate stretch, symmetric COO stretch, and OCO deformation modes [1]. In addition, XPS data showed a single sharp O ls line 113 [2] and a RAIRS study [5] of formate on the Cu(110) surface showed only a single COO stretch mode — although close inspection of the spectra suggests the possibility of a low-frequency shoulder. An apparent doublet in the CH stretch region was attributed to a combination band [5]. Figure 5.2 shows the average of four high-resolution spectra measured with the bolometer in the Cu-formate stretch region. There is clear evidence for a doublet. (Each of the individual bolometer spectra also showed evidence for a doublet; the lower resolution photoconductor scans, however, were less consistent.) The lines show the best fit to a sum of two Lorentzians after removing a linear baseline. The best fit values for the resonance frequencies of the two peaks are 365 and 380 cm'l, with linewidths of 18 and 13 cm‘l, respectively, and effective charges of 0.24e and 0.30e, arbitrarily assuming equal concentrations of 0.5 ML for each species. Both of these frequencies are considerably higher than observed in the EELS studies, which found the peak frequency to be 320-340 cm'l. The EELS experiments, however, did not use oxygen predosing, so the formate coverage was probably lower than in this experiment and some of the structural studies [3,4,10]. This coverage difference — or other effects arising from the oxygen predose — may account for the difference in frequency. The presence of a doublet suggests the possibility that multiple binding sites are occupied. The two sites most consistent with experimental and theoretical results are the "short-bridge" and "cross-bridge" sites [13]. We are l. ' 4'." ’"J 114 0.03 r .“ 0.00 £:;=u-..\._‘ '-‘__._.:Q.‘::;2?‘=’J_:_f_:.m " .07.‘\\\o . I” ”Z / a.“ ' . gs -0.03 - .-\‘\.\ 1/ - r V ' \ ’1 E -0.06 - ‘\° 0.] .. -009 — \3 g} _. E , . ‘1 . .. . ‘0.1 2 " J .4 310 330 350 370 390 410 430 450 Frequency ( cm:1 ) Figure 5.2 High resolution (6 cm'l) RAIR spectrum of formate on Cu(100), showing the bands assigned to the Cu-formate vibration. The lines show the best fit to a sum of two Lorentzians plus a linear baseline. 115 not aware of any calculation of the vibrational frequencies of formate for these sites on Cu(100); however Ushio et al. [30] have calculated the frequencies for two sites on Ni(110). They find a frequency of 368 cm'1 for the short-bridge site and 252 cm‘1 for the atop site. To the extent that these results can be carried over to Cu(100), they tend to confirm the identification of the binding site(s) as bridging, and to suggest that the Cu-formate stretch vibration is rather sensitive to the binding site. The small frequency difference (15 cm'l) between our observed peaks therefore suggests sites that are chemically, and perhaps geometrically, very similar. This similarity could account for the observation of only a single O ls line in XPS. Without additional experimental or theoretical information, we cannot determine the nature of the two sites. One possibility is a mixture of short-bridge and cross-bridge sites. Another is that all the sites have essentially the same geometry, but the long-range structure of the overlayer generates chemical differences due, for example, to antiphase boundaries between locally ordered regions. 5.4 Summary We have investigated the chemisorption of formate on an oxygen- predosed Cu(100) surface with RAIRS, using synchrotron radiation. No change in the broadband IR reflectance of the surface is detected upon adsorption of formate species. This suggests that formate scatters conduction electrons much more weakly than oxygen, CO, or hydrogen; the difference 116 may be due to formate's lack of electronic states near the Fermi level. Two low-frequency vibrational modes are observed, attributable to the Cu-formate stretch vibration. This suggests a mixture of two different binding sites. There is a clear need for additional RAIRS experiments on this system, to investigate the higher frequency internal vibrations, the effects of varying the coverage and the temperature, and the influence of the oxygen predose. Theoretical efforts to calculate the vibrational frequencies associated with different binding sites would also make an important contribution. 117 References 1.’ 10. 11. KC. Lin, P. Dumas, and R.G. Tobin, J. Vac. Sci. Technol. A 13, 1579 (1995). BA. Sexton, Surf. Sci. 88, 319 (1979); RA. Taylor, P.B. Rasmussen, C.V. Ovesen, P. Stoltze and I. Chorkendorff, Surf. Sci. 261, 191 (1992). M. Bowker and R.]. Madix, Surf. Sci. 102, 542 (1981). D.A. Outka, RJ. Madix and]. Stéhr, Surf. Sci. 164, 235 (1985). I. Stéhr, RJ. Madix, D.A. Outka and U. Débler, Phys. Rev. Lett. 54, 1256 (1985). BE. Hayden, K. Prince, D.P. Woodruff and A.M. Bradshaw, Surf. Sci. 133, 589 (1983); Phys. Rev. Lett. 51, 475 (1983). A. Puschmann, J. Haase, M.D. Crapper, C.E. Riley and D. P. Woodruff, Phys. Rev. Lett. 54, 2250 (1985). MD. Crapper, C.E. Riley and DP. Woodruff, Surf. Sci. 184, 121 (1987). MD. Crapper, C.E. Riley and DP. Woodruff, Phys. Rev. Lett. 57, 2598 (1987). MD. Crapper, C.E. Riley, D.P. Woodruff, A. Puschmann and I. Haase, Surf. Sci. 171, 1 (1986). DR Woodruff, C.F. McConville, A.L.D. Kilcoyne, Th. Lindner, I. Somers, M. Surman, G. Paolucci and A.M. Bradshaw, Surf. Sci. 201, 228 (1988). Th. Linder, I. Somers, A.M. Bradshaw and GP. Williams, Surf. Sci. 185, 75 (1987). 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 118 P. Hofmann and D. Menzel, Surf. Sci. 191, 353 (1986). LS. Caputi, G. Chiarello, M.G. Lancelotti, G.A. Rizzi, M. Sambi and G. Granozzi, Surf. Sci. 291, L756 (1993). SP. Mehandru and AB. Anderson, Surf. Sci. 219, 68 (1989). ].A. Rodriguez and GT. Campbell, Surf. Sci. 183, 449 (1987). A. Wander and B.W. Holland, Surf. Sci. 199, L403 (1988). M. Casarin, G. Granozzi, M. Sambi, E. Tondello and A. Vittadini, Surf. Sci. 307-309, 95 (1994). CJ. Hirschmugl, G.P.Williams, F.M. Hoffmann and Y.]. Chabal, Phys. Rev. Lett. 65, 480 (1990). CJ. Hirschmugl, Y.]. Chabal, F.M. Hoffmann and GP. Williams, J. Vac. Sci. Technol. A 12, 2229 (1994). KC. Lin, R.G. Tobin and P. Dumas, Phys. Rev. B 49, 7273 (1994). KC. Lin, R.G.Tobin, P. Dumas, CJ. Hirschmugl and GP. Williams, Phys. Rev. B 48, 2791 (1993). GP. Williams, private communication. R.G. Tobin, Phys. Rev. B 45, 12 110 (1992). B.N.]. Persson, Phys. Rev. B 44, 3277 (1991). B.N.]. Persson and Al. Volokitin, J. Electron Spectrosc. Relat. Phenom. 64/65, 23 (1993); Surf. Sci. 310, 314 (1994). GP. Williams, Int. 1. Infrared Millimeter Waves 5, 529 (1984); Nucl. Instrum. Methods A 291, 8 (1990). 27. 28. 29. 30. 31. 32. 119 GP. Williams, C]. Hirschmugl, E.M. Kneedler, E.A. Sullivan, and DP. Siddons, Y.J. Chabal, F. Hoffmann and K.D. Moeller, Rev. Sci, Instrum. 60, 2176 (1989). W. Jacob, V. Dose, and A. Goldmann, Appl. Phys. A 41, 145 (1986). J. Rogozik, V. Dose, K.C. Prince, A.M. Bradshaw, P.S. Bagus, K. Hermann, and Ph. Avouris, Phys. Rev. B 32,4296 (1985). K.-D. Tsuei and PD. Johnson, Phys. Rev. B 45, 13 827 (1992). R. Courths, B. Cord, H. Wern, H. Salfeld, and S. Hiifner, Sold State Commun. 63, 619 (1987). J. Ushio, I. Papai, A. St-Amant and DR. Salahub, Surf. Sci. Lett. 262, L134 (1992). Chapter 6 Adsorbate-Induced Reflectance Change of a Metal: CO on Pt Ipresent an investigation of the coverage dependence of the infrared reflectance changes induced by CO adsorption on Pt(111) she near-surface conductivity of a metal by the adsorbates causes a decrease in the reflectanceurface at room temperature. The modification of t. The magnitude of the reflectance change AR/R, at room temperature and high frequencies, increases monotonically at low CO coverages, peaks at 0:033 ML, and then decreases toward saturation coverage. The monotonic dependence is explained by a model of conduction electron scattering from the randomly distributed adsorbates. Possible explanations for the nonmonotonic behavior include coherent scattering from a partially ordered overlayer or a coverage- dependent scattering cross section. The maximum reflectance change is 0.28%:t0.03%, a factor of three smaller than 1.1% observed for oxygen or CO on Cu(100). The discrepancy can be explained by the scaling factor of 1/6, the ratio of skin depth and electron mean free path. This chapter is largely based on a paper to be published in Chemical Physics [1’]. 6.1. Introduction CO on Pt(111) has been extensively studied due to the practical application to the catalytic oxidation of carbon monoxide. On Pt(111), CO is 120 121 adsorbed non-dissociatively with the carbon atom bonded to the surface in a linear (on top) or a bridged configuration depending on the coverage. The internal C=O and external Pt-C stretching modes for CO on Pt(111) are well studied by infrared spectroscopy [1-3] for the complexity of the coupling and energy transfer between vibrational and electronic states of the adsorbates and substrate. Even without strong coupling, adsorbates can affect the electron dynamics in the near surface region as impurities from which conduction electrons will scatter without conserving the parallel momentum. These scattering events cause an increase in the dc resistance since the adsorbate perturbs the electrical transport to a depth on the order of the electron mean- free-path I, typically ~100 A, even though the scattering potential is screened within a few A [4]. It was also recognized [5—8] that the diffuse scattering on the surface would reduce the infrared reflectance. The observation is difficult to achieve because the broadband reflectance measurements demand that the absolute intensity be held constant to an accuracy considerably better than 1% between the measurement on the clean surface and that on the adsorbate- covered surface. Vibrational measurements on the other hand, are relatively insensitive to change in signal level as long as the background remains relatively smooth. Recent experiments [9-21] and theoretical developments [22-26] have revived interest in these scattering effects as they relate to adsorbate-substrate dynamics. A number of careful infrared studies on single crystals in 122 ultrahigh vacuum (UHV) have revealed adsorbate-induced decreases (~1%) in the broadband reflectance of metal surfaces that I have discussed in previous chapters. In this chapter, I mainly report the coverage-dependent adsorbate-induced reflectance changes for CO on Pt(111) and discuss some possible explanations for the observations. 6.2 Experimental The single crystal platinum sample was purchased from Aremco, Inc. The original sample was rectangular, with dimensions 3 cm wide x1 cm high x0.1cm thick. We cut off the sample edge at one corner by 1.5 cm x 0.3 cm in a triangular shape due to the presence of a domain misoriented by 9° from the [111] direction. The sample was repolished mechanically with 0.5 micron diamond paste. The normal to the sample surface of the remaining crystal is oriented at 27° from the [111] direction, as determined by the back reflection Laue X-ray technique. The crystal was also examined with low energy electron diffraction (LEED). The LEED pattern shows a doublet that is from the ordered step structure. The intra-doublet separation is inversely proportional to the distance between steps, so the terrace width can be obtained by comparing the doublet separation with a major spot separation. The terraces are 21-atoms wide as calculated from the doublet separation of the LEED pattern. The intensity distribution within the doublets can be obtained as a function of step height from kinematic analysis [27]. The calculated step height is 2.12 A. 123 From the terrace width and step height, the normal to the sample surface is at 273° from the [111] direction, which is in good agreement with Laue X-ray diffraction. We don't expect the 5 % density of step sites resulting from the miscut to affect our results; at the coverages studied the step sites should be saturated with CO, and the change in scattering rate due to the terrace should be independent of the presence of scattering centers on the step edges (Mathiessen's rule). The sample was mounted at the end of a stainless steel tube which can serve as a liquid nitrogen reservoir for cold fingers, in UHV at a base pressure S 1x10'10 torr. The sample was heated radiatively by a tungsten filament and the temperature was measured by two chromel-alumel thermocouples spot- welded to the edge of the sample. The sample could be closed with various gases by means of a leak valve and a multihole effusive array [28], which provides a flux enhancement of about 16 over background dosing. The primary contaminants in the crystal were carbon and calcium. Cleaning involved cycles of argon ion sputtering, annealing to 1170 K, exposure to oxygen at various elevated temperatures, and heating to 1360 K (to remove the atomic oxygen). After numerous cycles of sputtering and oxygen treatments, residual concentrations of C, Ca and O could be reduced to less than 3% of one monolayer (the noise level of ABS, i.e. 4% of the Pt 237 eV peak) and the integrated intensity of the oxygen TPD peak reached a maximum. 124 The spectroscopic system was the MSU instruments descirbed in chapter 2 [3]. A conventional silicon-carbide globar was used as the IR source, which was intensity modulated with a tuning fork chopper. The source housing, chopper, and source optics were cooled to liquid nitrogen temperature in high vacuum. The IR light was incident on the sample at 860 through a Csl window [29]. The reflected IR light passed through a second CsI window and focused on a home-built Czerny-Turner type infrared grating spectrometer housed in high vacuum and cooled to liquid nitrogen temperature to reduce the background photon flux. After the monochromator, the p-polarized component of the light is selected and detected by a Si-B photoconductor operated at liquid He temperature. The data taking procedure went as follows. First the sample was flashed to 975 K to clear the surface of a passivating layer of CO. The spectrometer was then set to a single frequency position and the IR reflected 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 6.1 shows the change in the reflected IR intensity versus time for a single measurement, normalized to the initial value, with a systematic linear baseline drift subtracted. The uncertainty was determined form the slope of the baseline drift prior to closing. This particular run shown exhibits 125 0.0005 "' .4 0.0000 '1 '1‘ '1‘ ..'.1H."Hl'lll‘..'."l - ~0.0005 - a % -0.0010 - a - 0.42:t0.03 ML CO/Pt(100) . -0.0015 - 2800 cm'l - 315 K n 1- . 1 ll 1 l I n. -0.0020 [ll “I”. “l ‘ . -0.0025 - -—>1 dosing 4_ .. 0 100 200 300 400 500 Time (s) Figure 6.1 Fractional change in reflected IR intensity from Pt(111) as a function of time at a frequency of 2800 cm'1 and room temperature. The Pt(111) sample was exposed to CO for 100 seconds at about the 200th second. Background pressure during dosing was 8x10’lo torr, while the pressure at the sample was increased by a factor of 16 due to the effusive array doser. A linear drift has been subtracted. 126 an extremely stable baseline. In most cases, the baseline drifts ~ 0.2% over the course of a run and the accuracy of the reflectance change measurement is limited. Usually, the drift was quite linear and AR / R can be measured to an accuracy of 0.02% or better. During this run the sample was closed with CO for 100 seconds beginning at about the 200th second at a background pressure of 8.0x10‘10 torr. The fractional change in the reflected IR intensity for this run was 0.00200i0.00002. It is of interest to compare the sensitivity of this home-built system for broadband reflectance measurements to that of the U4IR beamline at National Synchrotron Light Source of Brookhaven National Laboratory, where most of the previous studies have been carried out [12—16, 20]. Although the IR flux from the synchrotron radiation is greater than the traditional IR source in our system, electron beam movement and uncertainties in beam current limit the sensitivity to 0.2% [14, 20], about an order of magnitude above our level. It would be extremely difficult to investigate the reflectance change of the adsorption system such as CO on Pt(111), which have a maximum magnitude less than 0.3%. A significant drawback of our system is that we only measure a single frequency in each run, whereas the FTIR spectrometer at NSLS obtains the entire spectrum at once. The reflectance changes are measured for various coverages at 2500 and -1 . 2800 cm where the reflectance changes are predicted to reach an asymptotic 127 value within the scattering model [24,25], i.e., AR/ R becomes frequency- independent for a)>> 1'1, (01. As shown in Table 6.1, 1'1 corresponds to at most 970 cm'1 while (01 is considerately lower. Also the absoprtion due to the vibrational resonances is less than 0.003% from the Lorentzian oscillator approximation. The closest resonance peaks are the internal stretch of CO on Pt at ~1850 ch for bridge-bonded CO and at ~2100 cm.1 for atop-bonded CO [1.3]- The CO coverage was determined with reference to the C=O stretching frequency of atop CO. When measuring AR / R at a certain frequency, we first measured the clean-surface reflectance from 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 CO stretching vibration. The saturation coverage at room temperature was assumed to be 0.5 ML [2] and the rate of change of frequency with coverage was taken from Fig. 3. of Ref. [1], which is a compilation of published data. Uncertainty in the actual room temperature saturation coverage of our sample could introduce an overall shift in the coverage scale. 128 Table 6.1. Bulk material parameters for Pt. w,(cm'1) 011cm") 1 1cm") 1 (A) 5030 Opticala 41,500 133 558 91 380 [)(:b 77,000 372 968 79 207 aDrude parameters calculated from optical data [30]. bConduction electron density based on one electron per atom; scattering time calculated from DC resistivity measurements [31]. 6.3 Results and Discussion: Coverage Dependence In Fig. 6.2 (a), the fractional reflectance change AR/ R is shown as a function of CO coverage for CO on Pt(111) at 315 K at both 2500 and 2800 cm”1 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. The overall coverage- dependence is similar for O on Cu(100) (see chapter 4). The magnitude of the reflectance in that case increased monotonically up to 0.25 ML, but decreased by about 20% when the coverage increase dto 0.35ML. 129 Figure 6.2. (a) Fractional reflectance change AR/R as a function of coverage for CO on Pt(111) at 315 K. Data shown were taken at two IR frequencies, 2500 and 2800 cm'l. (b) Work function change for CO on Pt(111) from Refs. [34] (triangles) and [37] (circles). AR/R A(eV) 130 0.0000 . , . T . j , _l -0.0005 — _._ 2800 cm" + 2500 cm" -0.0010 '- i—L—i -0.001 5 '- + t—l——-1 5—1 -00020 - HEL- ” l—iF—Fj a 1 I 1 I - 1— H 0.0025 Egg: -00030 — (a) i 1 I 1 I 1 I 1 ‘ I ' I ' T T I ' I 0.00 ~ ‘ -0.04 - ‘ -0.08 r A -0.12 - O -0.16 - . _0.20 1 I 1 I 1 I 1 I 1 I 0.0 0.1 0.2 0.3 0.4 0.5 CO coverage (ML) 131 For a dilute, disordered overlayer, the scatterers are independent and the scattering of electrons will be incoherent-- the scattering rates from different adsorbates will simply add. If the scattering cross section per adsorbate is constant, the resulting reflectance change will vary linearly with adsorbate coverage. A linear dependence is observed at low coverages [10, 12,13,15, 16, 18, 19]. For CO on copper surfaces, the linear dependence persists up to saturation coverage [13, 16]. For other systems, however, nonmonotonic dependence at higher coverages have been observed. Riffe et al. [9, 10] measured the attenuation of surface electromagnetic waves (SEW) for N2, 0;, CO, H2, and D; on W(100). The magnitude of the SEW attenuation coefficient a increased monotonically with coverage for N2, while for both O2 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 study of H on W, Riffe and Sievers correlated the coverage-dependence of the reflectance change with ordering of the overalyer, at a frequency where the reflectance change was dominated by scattering [17] rather than electronic state effects [11]. At least two explanations can be offered for a nonmonotonic coverage- dependence. The scattering cross section per adsorbate could vary with coverage, either because different binding sites are being occupied or due to coverage-dependent changes in the electronic structure of the adsorbates. 132 Alternatively, ordering of the overlayer could reduce 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 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 the scattering amplitudes are in any case different, the reflectance need not return to its clean surface value [16]. On the other hand, it can be expected to be different from the reflectance from a disordered overlayer of the same surface density. Indeed, Riffe et al. interpreted their observations in terms of ordering, and found correlations between the observed coverage- dependence and adlayer patterns observed in separate low energy electron diffraction (LEED) experiments [32]. For O on Cu(100) surface, although long range order observable with LED could not achieved under our experimental conditions, the reflectance drops at a coverage where formation of an ordered overlayer structure is suggested by detailed studies of the overlayer structures [15]. It is interesting to note that 0.33ML is the coverage at which the CO overlayer forms a (43 Xw/3)R30° ordered structure on Pt(111) at 100 K, as observed by LEED [32]. At 0.5 ML the overlayer forms a c(4 x2) ordered structure [32]. It is tempting to attribute the decrease in AR/R to ordering. Since long-range order cannot be achieved at room temperature, however, 133 the effect would have to be assigned to partial, short-range ordering, for which there is no direct experimental evidence. Another possible explanation for the nonmonotonic coverage- dependence is a coverage-dependent scattering cross section caused by changes in the chemical bonding of the CO. It is well know that the CO occupies both atop and bridging sites on Pt(111) and that the relative populations of the two sites are coverage-dependent [34,35]. These two species have distinctly different electronic state and dipole moments [35]. So it is plausible that they might have different scattering cross sections for conduction electrons. Site exchange alone, however, can not account for the drop in AR/R seen in Fig. 3.2 (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: AR ——R—(0.5ML) — 0.25 %(0.33ML) 0-33 = 0.76 Eq. 6.1 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 0.5 ML [34, 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 measurements of the work function 134 change Ad) [34, 36, 37], which show a coverage dependence strikingly similar to that of AR / R, as shown in fig. 6.2 (b) where the work function data of Ref. [34,37] are reproduced. (It should be noted, however, that the coverage scale for Ref. [34] was set by assuming that the minimum value of Ad) occurred at 0.33 ML; in Ref. [37] the coverages were determined directly from TPD and LEED). The rapid increase in Ad) between 0.33 and 0.4 or 0.5 ML is attributed 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 Fig. 6.2 (a), but the ordering hypothesis cannot be ruled out. The magnitude of the fractional reflectance change AR/ R for p- polarized light on the surface is strongly dependent on the substrate properties and can be written [24]: £ _ 301:0" __P_) (R )_— __f(w/w1'l/6)'4ccose Eq.6.2 where p is the specularity parameter, which is a phenomenological parameter to describe the specularity of the scattering events of the conduction electrons on the surface ,0 is the angle of incident of the light and Up is the Fermi velocity in the metal, which within the Drude model depends only on n. The universal function for the frequency dependence fan/(1)1 ,1/5) depends on the frequency a) relative to 01:13:60,, / c and the ratio of the electron 135 mean-free—path I to the classical skin depth fizc/a)P. Near the characteristic frequency (01 an electron travels a distance comparable to the skin depth during one cycle of the magnetic field; (1); therefore specifies the frequency range below which nonlocal electrodynamic effects are important [22-24]. As I have pointed out the parameters that enter into f do not involve properties of the adsorbate; the frequency dependence depends only on the metal, while the specularity depends on the adsorbate-induced surface scattering. Under the same influence of the adsorbates, i.e. same p values, different metals exhibit different 1/5 values which scale the function f and affect the magnitude of the reflectance changes at high frequency limit. The calculated curves for these characteristic function are plotted in Fig. 6.3. from Ref. [24]. For copper, 1/8 is about 4.0 and the maximum reflectance change observed is ~1.1%. For Pt, 1/5 is about 0.2 and the maximum reflectance change observed is ~0.3%. That explains the large discrepancy of the reflectance changes shown between copper and Pt. This is also a characteristic due to the anormalous skin effect for metals. 136 E 2.5 1 1 r 5 0 e ./ .1: 2.0 — 0/ -1 “é / «3 Pt : 0 Cu E 1.5 " “ g}, 1.0 — ° - :1 C6 .5: 0 g 0.5 - - S: «3 O 8 <1) 31:, 0.0 . I . 9‘ 10'2 10'1 100 101 102 Scaling factor 1/5 at a frequency of 60/61), = 4 Figure 6.3 The reflectance change AR/R vs. US, the ratio between the bulk electron mean free path and the skin depth, at a frequency 60/(11)l = 4 where the reflectance change magnitude approach to an asymptotic value for the universal frequency-dependent function. (reproduced from Ref. [24]) 137 6.4 Summary We have measured the adsorbate-induced change in the broadband IR reflectance change of Pt(111) as a function of CO coverage. The IR reflectance decreases when CO adsorbs on the Pt(111) surface. After increasing with increasing coverage for low coverages, the magnitude of the reflectance change peaks at a coverage of about 0 = 0.33 ML, and decreases toward saturation. The presence of a peak in the coverage dependence could be attributable to ordering in the overlayer or changes in the electronic structure of CO-Pt bonding. The observed reflectance change for CO/Pt is a factor of three smaller than that measured for O/Cu. The discrepancy can be qualitatively explained by the scaling factor 1/5. 138 References 1.’ 10. 11. 12. 13. 14. DE. Kuhl, K.C. Lin, Chilhee Chung, J.S. Luo, Hong Wang, and R.G. Tobin, Chemical Physics, in press. J.S. Luo, R.G. Tobin, and D.K. Lambert, Chem. Phys. Lett. 204, 445 (1993). 1.]. Malik and M. Trenary, Surf. Sci. 214, L237 (1989); LR Sutcu, J.L. Wragg and H.W. White, Phys. Rev. B 41, 8164 (1990). C. Chung, Dr. thesis, MSU, 1993, references there in. K. Fuchs, Proc. Cambridge Phil. Soc. 34, 100 (1938). T. Holstein, Phys. Rev. 88, 1427 (1952). G. E. H. Reuter and E. H. Sondheimer, Proc. R. Soc. London, Ser. A 195, 336 (1948). R. B. Dingle, Physica 19, 311 (1953) R. B. Dingle, Physica 19, 729 (1953). D. M. Riffe, L. M. Hanssen, and A. J. Sievers, Phys. Rev. B 34, 692 (1986). D. M. Riffe, L. M. Hanssen, and A. J. Sievers, Surf. Sci. 176, 679 (1986). J. E. Reutt, Y. J. Chabal, and S. B. Christman, Phys. Rev. B 38, 3112 (1988. C. J. Hirschmugl, G. P. Williams, F. M. Hoffmarm, Y. J. Chabal, Phys. Rev. Let. 65, 480 (1990). C. J. Hirschmugl, Y. J. Chabal, F. M. Hoffmann, G. P. Williams, J. Vac. Sci. Technol. A 12(4), 2229, (1994). K. C. Lin, R. G. Tobin, P. Dumas, C. J. Hirschmugl, and G. P. Williams, Phys. Rev. B 48, 2791 (1993). 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28 29. 139 K. C. Lin, R. G. Tobin, and P. Dumas, Phys. Rev. B 49, 17273 (1994). C. J. Hirschmugl, G. P. Williams, B. N. J. Persson, A. I. Volokitin, Surf. Sci. Let. 317, L1141 (1994).22. D. M. Riffe and A. J. Sievers, Surf. Sci. 210, L215 (1989). E. Borguet, J. Dvorak, and H. L. Dai, Laser Techniques for Surface Science, 2125, 12 (1994). M.Hein and D. Schumacher, J. Physics D Applied Physics, 29, 1937 (1995). C.L.A. Lamont, B.N.J. Persson and GP. Williams,Chem. Phys. Lett., 24, 429 (1995). J. Krirn, D.H. Solina and R. Chiarello, Phys. Rev. Lett. 66, 181 (1991). B.N.J. Persson, Chem. Phys. Lett. 197, 7 (1992). B. N. J. Persson and A. I. Volokitin, J. Electron Spectrosc. Relat. Phenom. 64/65, 23 (1993). B. N. J. Persson and A. I. Volokitin, Surf. Sci. 310, 314 (1994). B. N. J. Persson, Phys. Rev. B 44, 3277 (1991). J. B. Sokoloff, Pys. Rev. B, to be pulished. M.B. Webb and MG. Lagally, in Vol. 28 of Solid State Physics, edited by H. Ehrenreich, F. Seitz and D. Turnbull, p.301 ( Academic, New York, 1973). D. E. Kuhl and R. G. Tobin, Rev. Sci. Inst. 66, 3016 (1995). R. G. Tobin, C. Chung, and J. S. Luo, J. Vac. Sci. Technol. A 12, 264 (1994). 30. 31. 32. 33. 35. 36. 37. 140 M. A. Ordal, R. J. Bell, R. W. A. Jr., L. L. Long and M. R. Querry, Appl. Opt. 24, 4492 (1985). J. Bass and K. H. Fischer, in Landolt-Bornstein Numerical Data and Functional Relationships in Science and Technology, 15a, edited by K. H. Hellwege (New York, 1982), p. 63. D. A. King and G. Thomas, Surf. Sci. 92, 201 (1980). H. Steininger, S. Lehwald, and H. Ibach, Surf. Sci. 123, 264 (1982). P. R. Norton, J. W. Goodale, and E. B. Selkirk, Surf. Sci. 83, 189 (1979). W. D. Mieher, L. J. Whitman, and W. Ho, J. Chem. Phys. 91, 3328 (1989). K. Horn and J. Pritchard, J. Phys. (Paris), Colloq. C4, Suppl. No. 10, 38, 164 (1977). G. Ertl, M. Neumann, and K. M. Streit, Surf. Sci. 64, 393 (1977). Chapter 7 Conclusion In this work, Ihave studied the adsorbate-induced reflectance changes on metal surfaces. Experimental results support the scattering model that suggests the reduction in reflectance and increases in surface resistivity are caused by conduction electron scattering. While the frequency-dependence of reflectance change reveals the conduction electron properties, the magnitude of the reflectance change provides information on the adsorbate-substrate interactions. The direct relation between reflectance change and resistivity change for O and CO on Cu(111) surfaces is proven within satisfaction by comparing the measured reflectance change and published resistivity change on thin metallic films under an assumption that the change in electron relaxation time is much larger than the change in conduction electron density. For CD on Ni, the reflectance change is much smaller than expected from resisitivity measurements, which can not explained by the scaling factor that implies the interaction between probing light and conduction electron and is the ratio of the classical skin depth to the electron mean free path. The discrepancy may result from different experimental condition and is subject of further investigation: that is the simultaneous optical and resistance measurements on well-defined surfaces. 141 142 The adsorbate-induced reflectance change is also characteristic of the interaction between conduction electrons and the incident electric field. The frequency dependence of the reflectance change shows the local-nonlocal transition, which is predicted by calculating the induced current due to the probing light from the distribution function of conduction electrons under the influence of the diffuse scattering from adsorbates. The characteristic frequency that determines the roll-off of the reflectance change depends only on the conduction electron properties. The magnitude of the reflectance change indicates the strength the interaction, i.e., scattering, between adsorbate and substrate and can be expressed through the specularity parameter. For oxygen on Cu(100) at coverages below 0.25 ML, the reflectance changes are well fitted by the universal form. At higher coverage, the frequency dependence can no longer be fitted by the theory and the magnitude of the reflectance changes and coverages are no longer in a monotonic relation. Possible explanations are: the formation of an ordered overlayer formation and the decrease of the effective scattering cross section. Upon exposure of an oxygen-predosed copper surface to formate, the reflectance returns to the clean surface value; there is no reflectance change detected, unlike H, O, CO on Cu surface. The explanation of the absence of the reflectance change is lack of an adsorbate-derived surface state since the direct interaction between adsorbate and conduction electron requires overlapping of surface induced electronic states and the Fermi level. 143 A nonmonotonic coverage dependence of the reflectance change for CO on Pt(111) is also observed. The magnitude of the reflectance change AR/R, at room temperature and high frequencies, increases monotonically at low CO coverages, peaks at 020.33 ML, and then decreases toward saturation coverage. The reduction of the reflectance change occurs when the ordered overlayer starts to form and the work function starts to decrease. However, we cannot really distinguish which effect dominates the reduction of the diffuse scattering. My study shows the scattering theory well explains the origin of the broadband reflectance changes and this information can provide a new prospect of surface IR study. We learn more about the conduction electron properties in the near surface region and how the adsorbate interacts with substrate that is exactly what surface scientists are trying to do in the last thirty years. Appendix: The intensity of an energy loss peak divided by the elastic intensity is (for the case of specular reflection): Iv 47We 2 2 N s — = ———e (E , a) 10 11250 W cosaf 0 IV: measured vibrational mode peak intensity IO: elastic peak intensity me: electron mass h: Planck’s constant N 5: number of dipoles per cm2 E0: primary electron ( incident electron energy) 0:: angle of incident 11V: effective dipole moment h #v = e *2 2——— mama m #2 effective mass of the oscillator a)”: vibration frequency of the oscillator in wave numbers e*: effective charge of the oscillator f(E0,a) = (sinz— 2cos2 a)Y +(sin2— 2cos2 a)lnX 144 145 __ 6% . -eiwé 012+90’ 03 61: half angle of the aperture of the detector 90: beam width 3‘1 2130 For uncoupled. oscillators: 2 AR 7: e* 2 1 sin261+rp I—dv= —-7 — —NS 2 R c e m” c036 Ir I P 0: incident angle of incoming light rpt reflectivity constant given by the Fresnel equations for p-polarized light The correlation between measured IR absorption intensity and the intensity of EEL signal is linked through dynamic electron charge. References 1. S. Anderson and J.W. Davenport, Solid State Commun. 28, 677 (1978); IS. Luo, Reading Notes, 1990. 2. H. Ibach, Surf. Sci. 66, 56 (1977); A.M. Baro, H. Ibach, and H.D. Bruchman, Surf. Sci. 88, 384 (1979). 3. C. Chung, Dr. thesis, MSU, 1993. nICHIan STATE UNIV. LIBRQRIES lllllllllllllllllllllllllllllllllllllllllllllll 31293013908383