. ... 2: .. 23.1..» ft? pl .2941. .: r: .10 i. r 7“. 3 1W - dark iguana .0 ’31: ‘1': «I .. L ‘1‘; Us. . {.1714 < \ l . ‘ V . . $§ I. fl‘v.txt A...- v. 30; o .3. wVvI-dl. H‘|.|.rj.V...L-.. a; It'bl'. | THESIS ERSIIBTYL BRARIES IHII IIHI l II HHIWHIWlllltilHiU 3 129300 This is to certify that the dissertation entitled DEVELOPMENT OF A NEW FAR-INFRARED REFLECTION- ABSORPTION SPECTROMETER AND MEASUREMENTS OF DIOXYGEN ON THE Pt(111) SURFACE presented by Chilhee Chung has been accepted towards fulfillment of the requirements for Ph.D. , Physics degree in Q7 774 flMajor professor Date July 22, 1993 MSU i: an Affirmative Action/Equal Opportunity Institution 0-12771 LIBRARY Michigan State University I PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. ‘ DATE DUE DATE DUE DATE DUE r j: ‘J‘ 7 T f TA _1 zfi MSU Is An Affirmative Action/Equal Opportunity Institution cWUnS-o.| DEVELOPMENT OF A NEW FAR-INFRARED REFLECTION- ABSORPTION SPECTROMETER AND MEASUREMENTS OF DIOXY GEN ON THE Pt(111) SURFACE By Chilhee Chung A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY . Department of Physics and Astronomy and Center for Fundamental Materials Research 1993 ABSTRACT DEVELOPMENT OF A NEW FAR-INFRARED REFLECTION - ABSORPTION SPECTROMETER AND MEASUREMENTS OF DIOXYGEN ON THE Pt(111) SURFACE By Chilhee Chung Infrared spectroscopy has long been widely used for examining adsorbates on solid surfaces, but until recently its use was largely confined to high frequency modes (above ~100013Ed), mostly internal modes of molecular adsorbates. In recent.years several groups have successfully measured the adsorbate-substrate vibrations of molecular adsorbates, and in ‘very limited cases atomic adsorbates, which lare usually at low frequencies below 1000 cmfl. Vibrational spectroscopic measurements of such low frequency modes with high resolution, high sensitivity and excellent baseline substraction will likely play an important role in the understanding of adsorption, surface dynamics, and other surface phenomena. There are only a limited number of surface’ infrared spectroscopic instruments that can be operated in the low frequency region with those capabilities. We have built a new far-infrared reflection-absorption spectroscopy system designed specifically for the frequency region 350 - 1000 cm"1 (usable up to 3500 cm“) . It incorporates a liquid-nitrogen—cooled grating spectrometer and other cooled optics to minimize noise due to ambient background radiation, a highly sensitive SizB photoconductive detector and a UHV sample chamber equipped with LEED, Auger and TPD. The instrument currently achieves a level of sensitivity necessary for the study of the low frequency modes, ~2 x 10“ l/fl'E at ~850 cm'l. We have measured the 0-0 stretch mode of molecularly adsorbed 02 on Pt(111) at ~875 cm'1 and the molecule-substrate vibration of CO on Pt(111) at ~466 curl. The dominant noise source is non-reproducibility from one scan to the next. An improvement of the N/S ratio by a factor of 4 is expected by improving the scan-to-scan reproducibility. As a first experiment, the isotope effect on the linewidth of the 0-0 stretch mode will be measured. The shifts in the linewidth are expected to be 1.2 cm"1 if inhomogeneous broadening is dominant, or 12.5 cm"1 if electron-hole pair broadening is dominant. With the resolution, N/S ratio and baseline subtraction performance of our new system, those shifts can be measured. We have measured the linewidth for 1‘02 on Pt(111) , 21.6 cm-1 and will measure the linewidth for 1802 to sort out the line-broadening mechani sm . To My Wife and Son iv ACKNOWLEDGMENTS I would like to thank my thesis adviser, Professor Roger G. Tobin, for his support, insightful and patient guidance, and encouragement throughout this work. He has taught me how to design, construct and evaluate the apparatus, and how to do surface experiments. Without his patient advising, this dissertation could not have been finished. I am grateful to many other people at MSU, in particular, the physics department machine shop staffs, Tom Palazzolo, Jim Muns and Tom Hudson, for their help in building the instrument, Peggy Andrews for her patient handling of many administrational work, and other members of our research group, Kathy Lin, HongflWang and.J.S. Luo for their help and collaboration. I would like to give special thanks to my wife and son, my mother, and parents-in—law for their support and.encouragement. I also would like to thank many friends at Sansung for their support and encouragement. This work is financially supported in part by the Center for Fundamental Materials Research, by the Petroleum Research Fund, administered by the American Chemical Society, and by the National Science Foundation under Grant No. DER-8815616. TABLE OF CONTENTS List of Tables ................................................. List of Figures ................................................ Introduction .................................................... . Scientific Background and Objectives .............................. 1.1 Line broadening mechanism ................................. 1.1.1 Lifetime broadening ................................. 1.1.2 Dephasing ........................................... 1.1.3 Inhomogeneous broadening ............................ 1.2 Experimental determination of the line broadening mechanism 1.2.1 Experimental methods ................................ 1.2.2 Summary of previous experiments ..................... 1.3 0—0 stretch.mode of 02 on Pt(111) .......................... 1.4 Summary .................................................... References .................................................... . Feasibility of the IRAS Experiment ................................ 2.1 Comparison of IRAS with other vibrational spectroscopies .. 2.2 Surface selection rule on metal surfaces .................. 2.3 Incidence angle dependence of the surface signal .......... 2.4 Requirements for an IRAS system ........................... 2.5 Calculation of photon noise ............................... vi xi 10 10 ll 12 15 18 22 24 28 28 30 32 36 37 2.6 General considerations of the IRAS instrument ............. 2.6.1 Radiation source .................................... 2.6.2 Spectrometer ........................................ 2.6.3 Detector ............................................ 2.6.4 Modulation .......................................... 2.7 Design of the IRAS system ................................. 2.7.1 Angle of incidence and throughput ................... 2.7.2 Spectrometer ........................................ 2.7.3 Minimizing ambient radiation ........................ 2.7.4 Detector ............................................ 2.8 Summary ................................................... References .................................................... . Description of the Apparatus ....................... _ ............... 3.1 Radiation source in a liquid-nitrogen cryostat ............ 3.1.1 Cryostat ............................................ 3.1.2 Globar source in a copper housing ................... 3.1.3 Ellipsoidal and plane mirrors ....................... 3.1.4 Tuning fork chopper.................................. 3.2 Transfer optics ........................................... 3.3 UHV system ................................................ 3.3.1 Surface analysis tools .............................. 3.3.2 Sample mount ........................................ 3.3.3 Infrared window ..................................... 3.3.4 Gas doser ........................................... 3.4 Grating Spectrometer ...................................... 3.4.1 Cryostat and structure .............................. vii 41 41 43 45 46 48 49 50 52 54 SS 56 68 68 69 70 70 71 72 73 75 76 77 77 79 79 3.4.2 Optical 3.4. 3.4. 3.4. 3.4.3 Grating 2.1 2.2 2.6 components .................................. Entrance slit ............................... Folding mirrors ............................. Paraboloidal mirrors ........................ Diffraction gratings ........................ Exit slit and filters ....................... 3.5 Cooled Transfer Optics .................................... 3.6 Detector OOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO 3.7 Data Acquisition .......................................... 3.8 Optical Alignment ........................................ References 3.8.1 Optical alignment of the spectrometer ................ 3.8.2 Optical alignment of the whole system ................ . System Performance ............................................ 4.1 Calibration of spectral frequency ......................... 4.2 Spectrometer resolution ................................... 4.3 4.4 4.5 4.6 Detector calibration ...................................... Optical efficiency of the system .......................... Calculated noises ......................................... 4.5.1 Noise of the detector circuit ....................... 4.5.2 Photon noise ........................................ Measured noises ........................................... 4.6.1 Noise from the detector circuit ..................... 4.6.2 Photon noise ........................................ viii 83 83 83 84 85 86 86 86 89 89 91 92 92 94 95 114 114 116 117 124 126 127 130 132 132 132 5. 4.6.3 Behavior of system noise ............................ 4.7 Improving noise-to—signal ratio ............................ 4.8 Comparison of our IRAS system and others .................. 4.9 Summary ................................................... References Measurements on 02/Pt(1 1 1) and CO/Pt(111) ......................... 5.1 Sample preparation and characterization ................... 5.2 Measurements of vibrational spectra for CO on Pt(111) ..... 5.3 Measurements of 0-0 stretching band of Ozton Pt(111) ....... References Conclusion Reference OOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO OOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO OOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO ix 134 136 139 141 143 156 156 158 160 167 178 180 2.4.1 3.4.1 4.3.1 4.5.1 4.8.1 5.1 LIST OF TABLES Peak IR intensities and frequencies for the adsorbate— substrate and internal vibrations for a number of molecular and atomic adsorbates on metal surfaces. External modes have weaker signals and lower frequencies, in general. Specifications of the diffraction gratings. Detector Calibration. The responsiVity of the detector has been measured for three different temperatures, using the calibration setup shown in Figure 4.3.1. The efficiency of the detector optics is included in the responsivity and the amplifier gain A. used is 10. Summary of calculated and.measured noises. See sections 4.5 and 4.6 for details. Comparison of IRAS systems. No groups, BNL group and Bermudez group, are using brighter sources to improve the signal intensity; while others are using 15 kcal/mole. Many surface scientists investigate the structural. dynamical. and electrical properties of the chemisorbed layer to study gas-surface interactions and many surface chemical reactions. Among many experimental techniques used to investigate surface properties, surface vibrational spectroscopy (SVS) has been known as an important tool because it has the capability to determine the chemical nature of adsorbates and surfaces that are not readily available by other techniques such as Auger electron spectroscopy. and ultraviolet or X-ray photoelectron spectroscopies (UPS or XPS). While electron and photoelectron spectroscopy are used to study electronic properties of the .core and valence states of adsorbates. surface vibrational spectroscopy can be used to identify the types of adsorbed atoms or molecules on a surface, and to determine bonding sites and adsorbate geometry. Bond strengths and distances of adsorbates can be estimated from the vibrational frequencies. Furthermore. information about surface dynamics can be obtained directly from time-resolved laser measurements or indirectly from line shape and linewidth analysis of SVS. There are many SVS techniques available; surface, infrared spectroscopy. electron energy loss spectroscopy (EELS). surface enhanced Raman spectroscopy. inelastic tunneling spectroscopy. and inelastic atom 4 or neutron beam scattering. Surface infrared spectroscopy can be further classified; infrared reflection-absorption spectroscopy (IRAS). infrared emission spectroscopy (IRES), surface electromagnetic wave spectroscopy (SEWS). and. electroreflectance ‘vibrational spectroscopy (EVS). among others. Since each SVS technique has its own advantages and limitations. the investigation of a particular surface vibrational phenomenon usually requires a specific SVS technique. In certain cases. it is also necessary to use more than one technique to perform an experiment; the combination of IRAS and EVS is used for a direct measurement of the Stark tuning rate for adsorbed molecules on surfaces. IRAS and EELS are among the most popular and successful techniques (see section 2.1). partly due to high sensitivity and versatility. The energy transfer between the adsorbed molecule and the substrate plays an important role in such surface dynamical phenomena as adsorption, desorption. surface diffusion, and surface chemical reactions. Therefore. the mechanisms and rates of the energy transfer have been among the most intensively studied subjects. both experimentally and theoretically, in surface dynamics [4-15]. Nevertheless. the understanding is far from clear for most adsorbate modes. particularly at low vibrational frequencies (below 1000 cm'l), partly due to the limited experimental data available. (There is no clear threshold for the term low frequency -- there are some experiments below 1000 cm’l. though not a lot. and only a handful below 800 cm".) One of the experimental approaches is. direct time-resolved measurements of the vibrational lifetime. A direct measurement would unambiguously identify the energy transfer channel and separate 5 homogeneous and inhomogeneous contributions to vibrational linewidths. The experiments are very difficult, and were at first limited to high- surface-area insulators. and long-lived modes (2 100 ps) [10]. Recent advances in the time-resolved laser technique have made it possible to directly measure the vibrational lifetimes for some adsorbates on well- characterized single crystal surfaces [ll-15]. An intense resonant picosecond infrared (pump) pulse creates an excited state population of an adsorbate vibrational mode. The subsequent relaxation of the excited vibrational mode is monitored by an infrared spectroscopy based on a second infrared (probe) pulse. For the second probe spectroscopy, both IRAS and surface IR-visible sum frequency generation (SFG) spectroscopy have been successfully used. The IRAS probe method has the advantage that the observed spectral line shapes may be compared to the vast available IRAS data. In the SFG spectroscopy method. the signal is generated through the nonlinear response of the surface at the sum frequency of the visible and infrared beam. and therefore it can be detected using the highly sensitive photomultiplier tube. The time-resolved measurements were performed for CO on Pt(111) [11.12] . cnas (methyl thiolate) on Ag(lll) [13], CO on Cu(100) [l4]. and H on Si(111) [15]. but they are still limited to high-frequency modes. mainly due to the lack of tunable far- infrared picosecond lasers. The other experimental approach is to examine the vibrational spectra of adsorbates on the surface. The frequencies. linewidths. and line shapes of the vibrational modes. and their dependences on coverage and temperature, contain information about the adsorbate-adsorbate and adsorbate-substrate interactions [5.8.9]. The linewidth. for example, can 6 be a measure of the finite lifetime of the vibrationally excited state, and therefore of the rate of energy transfer between the adsorbed molecule and the substrate. It is remarkable that information about a process on a picosecond time scale can be obtained through the static measurement of the vibrational linewidth. The linewidth. in general, can also be affected by other processes, such as dephasing and inhomogeneous broadening. In order to sort out these effects. it is essential to perfonm a precise line shape and linewidth measurement, which requires high sensitivity and resolution. and.the capability to vary such dynamical variables as temperature and coverage over a wide range. Electron energy loss spectroscopy (EELS) has good sensitivity, but does not have the resolution required for linewith studies. Surface infrared spectroscopy has good resolution and sensitivity, and it is regarded as nearly the only experimental technique adequate for studies of the linewidths and line shapes of adsorbate modes. The spectral region of conventional surface infrared spectroscopy has historically been limited to frequencies above 1000 cm'l. because of the low intensity of thermal IR sources and bright background radiation. Only a limited number of surface infrared spectroscopy instruments can be routinely operated in the low frequency region [17-20]. Most adsorbate-substrate vibrations have frequencies in the far-infrared region, below 1000 cm'l. and they play very important roles in dynamical processes. That is why we have built a new far- infrared spectroscopy system. which is specifically designed for the far- infrared region. below 1000 cm'l. This dissertation describes the design. construction. characterization. and preliminary results of this system. 1. Scientific background and objectives Surface infrared spectroscopy has proved to be one of the most useful experimental techniques for investigating the dynamics and interactions of adsorbed atoms and molecules on clean surfaces [4-7] . Due to such features as high resolution, strict selection rules, and quantitatively interpretable line shapes and intensities, it has been successfully used to obtain information about the chemical nature of adsorbates. adsorption sites. molecular orientation, adsorbate-adsorbate interactions]. adsorbate- substrate energy transfer. and electrostatic effects of surface defects. For the most part, however these investigations have been confined to high frequency modes, above 1000 cm'l. which are mostly the internal vibrational modes of molecular adsorbates. The low frequency adsorbate-substrate modes, usually below 1000 cm'l, have been historically inaccessible by conventional infrared spectroscopy. in part because the signals are usually weak and are interfered with by room temperature radiation [16]. They are. however, even more important than the high frequency modes to such surface phenomena as adsorption. desorption. surface diffusion. heterogeneous catalytic reactions. and other surface phenomena. In recent years. some adsorbate-substrate modes have been measured. but still they are not readily accessible. Here at MSU we have built a unique IRAS system specifically designed for studying such adsorbate-substrate modes. In this chapter. I discuss one of the subjects that will be investigated by the new instrument: the study of line-broadening mechanisms of molecular and atomic adsorbates on single crystal metal 8 surfaces. Of course. the instrument can and will be used for other studies. such as adsorbate-substrate vibrations of atomic and molecular adsorbates, alkali metal adsorption on metals. adsorption on semiconductors and broadband IR absorption of metal surfaces. which require high resolution, very high signal-to-noise ratio and excellent baseline subtraction. 1.1 Line broadening mechanism One of the most remarkable features of the infrared vibrational spectrum of a chemisorbed molecule is the large linewidth. The linewidth of the C-0 stretch vibration of chemisorbed CO. for example. is typically ~ 10 cm'l. some 1010 times greater than the natural linewidth of the same vibration in the gas phase [5]. The broadening processes can be largely classified as lifetime broadening. dephasing. and inhomogeneous broadening. 1.1.1 Lifetime broadening The vibrational energy of the adsorbed molecule could decay and 'result in lifetime broadening by exciting or creating such excitations as photons. phonons. adsorbate vibrational modes other than the one under study. and electron-hole pairs (for metal substrates). First, the decay process which generates a photon is entirely negligible. since the radiative lifetime of the C-0 stretch vibration is estimated to be on the order of 10"1 second, which is some 1010 times longer than the lifetime implied by the adsorbate linewidth [5]. Second. lifetime broadening by phonon emission has received a lot of attention at least for modes with vibrational frequencies less than two or three times the maximum substrate 9 phonon frequency [21-24.33,38]. Since typical maximum phonon frequencies are 200 - 500 cm*l[2]. this process is hardly significant for such high- frequency modes as C-O at ~2000 cm'1 on metal surfaces. but it has long been thought be at least potentially important for low frequency modes. Recently. however, Persson and Ryberg [25] have pointed out that the direct phonon contribution is negligible on metal surfaces, although it can still dominate on semiconductors [26] and insulators [27]. Third. the decay of a vibrational mode could occur through the excitation of other adsorbate vibrational modes, accompanied by the emission or absorption of one or more substrate phonons. This process is difficult to explore experimentally. since some vibrational modes cannot be measured with infrared spectroscopy due to the surface selection rule (see section 3.2) and others are even unknown. There is, however. a pretty clear example. Harris et a1. [13] have directly measured the vibrational relaxation time of the C-H stretch mode of CHQS on Ag(lll). using SFG spectroscopy. They have observed two component relaxation times. and have assigned them to intramolecular vibrational relaxation. proposing the C-H asymmetric stretch and/or bending modes as candidates for the intermediate vibrational mode. Lastly. on metallic substrates, electron-hole pair generation could be the dominant channel for vibrational decay [8,28- 32.42.54]. Since the number of accessible electron states increases linearly with the energy of the vibration. the decay rate for this process increases with vibrational frequency. It is therefore most likely to dominate the linewidth for high frequency modes that are well above the maximum substrate phonon frequency. Langreth [30] has argued. based on the idea of an adsorbate-induced resonance near the Fermi level. that the 10 electron-hole pair mechanism necessarily produces a characteristic, asymmetric line shape with a extended tail toward either side of the peak frequency depending on certain relevant parameters. Volokitin [32] has further predicted that a similar asymmetric line shape should be produced for a vibration parallel to the surface. as well as for a vibration perpendicular to the surface. if the line is broadened by the electron- hole pair mechanism. This argument is discussed more fully in section 5.3. 1.1.2 Dephasing In addition to lifetime broadening. dephasing can provide a homogeneous contribution to the linewidth [8]. The process of dephasing has been well known in other fields. such as nuclear magnetic resonance and optical spectra of impurity ions in crystals. Contrary to lifetime broadening, pure dephasing results in broadening the vibrational linewidth with no decay in amplitude. The dephasing process can be caused by an impulsive elastic collision of the vibrating molecule with a phonon [8.33], by a coupling to another adsorbate vibrational mode at lower frequency [32,34]. or by the interaction of the adsorbate vibration with electron-hole pairs [32,35]. 1.1.3 Inhomogeneous broadening The inhomogeneous contribution to the linewidth is believed to dominate in partial overlayers of adsorbates on metals since the narrowest linewidth is observed for the well ordered overlayer. In reality. some degree of inhomogeneous broadening is expected even for well ordered overlayers. since all real adsorbed layers have inhomogeneities due to finite substrate. finite domain size and domain boundaries. and a certain 11 amount of ‘point, defects and. contaminants that always exist on the substrate. These inhomogeneities disturb the local chemical environment to break the ideal periodicity of the well ordered overlayer, resulting in shifts of the vibrational frequencies of nearby adsorbates [5]. If long range dipole coupling is important. these effects could extend over a appreciable region of the surface. The extent of inhomogeneous broadening is not. however. completely known for the well ordered overlayer [8]. Since the inhomogeneity of a _real surface is often too complicated and is caused by many different factors. it is extremely difficult to analyze either the line shape or the linewidth even for a complete layer. It is often assumed that the line shape will be asymmetric. Recently Jacob et a1. [36] have made measurements for H on Si(lll) prepared by an etch in a buffered HF solution. a surface that closely approximates an ideal complete layer. They have measured an asymmetric line shape with a low frequency tail and a linewidth of 0.05 cm'l. which is apparently due to inhomogeneous broadening. and have shown that two major sources are responsible for the asymmetric inhomogeneous broadening: normal modes other than the in-phase normal modes and the domain size distribution. It is generally expected that the extent of inhomogeneous broadening will be sensitive to the number. nature. and distribution of defects and contaminants. Inhomogeneous broadening is therefore likely to depend on the preparation of the surface. 1.2 Experimental determination of the line broadening mechanism As discussed above the line broadening contains the combined effects of non-radiative vibrational decay (T1 process). dephasing (T2 process). 12 and inhomogeneous broadening. In general. it is difficult to determine experimentally the dominant broadening mechanism for a given vibrational mode. The extent of inhomogeneous broadening should be sorted out to estimate the homogeneous line broadening. This task is. however. very difficult to realize. It is essential to minimize the inhomogeneous effects and to work.on.well characterized systems for the careful study of linewidth and line shapes. The surface should be smooth. as free as possible of crystal defects and contaminants. as determined by X-ray diffraction and Auger electron spectroscopy (AES). and adsorbate overlayers should be well ordered and complete. as determined by low energy electron diffraction (LEED). Nevertheless. the possibility of inhomogeneous broadening always exists even for the well characterized system [5]. In this section. experimental methods to distinguish the line broadening mechanism and previous results are presented. 1.2.1 Experimental methods The homogeneous broadening could be isolated from the inhomogeneous contribution experimentally by careful.measurements of the line shape. and of the mass and temperature dependences of the linewidth. Each line broadening mechanism is expected. or predicted, to give characteristic features for the measurements. For studies of linewidth and line shapes. the measurements require‘very“high signal-to-noise ratio. high resolution, and a very flat or smooth baseline over an extended region. since the differences are very small and often subtle. First. homogeneously broadened lines are expected to be Lorentzian [8], if the line is symmetric. and are also predicted to be asymmetric if they are caused by the generation of electron-hole pairs in the substrate 13 [30]. Inhomogeneous lines are often highly asymmetric, and it is often assumed (though with little justification) that they will be Gaussian.[8], if they are symmetric. Second. the mass of the adsorbate can be' changed by isotopic substitution without affecting the chemical environment. The characteristic mass dependence for various line broadening mechanisms. however. is not clear and is controversial in some cases [5]. Let us define the quantity a - (a 1n 7/6 1n mg) to represent the mass dependence of the linewidth. where 1 is the linewidth. and m.Jr is the reduced mass of the adsorbate. The isotope dependence of electron-hole pair line-broadening has been predicted theoretically. Persson. et a1. [37] have predicted that electron-hole pair decay should have a - -1. Volokitin [32] has also predicted the same mass dependence of the linewidth. for vibrations both parallel and perpendicular to the surface. and has further predicted that the intensities at peak frequency have different mass dependences: mgqu _for perpendicular vibration and no mass-dependence for the parallel vibration. The isotope dependence of dephasing due to phonons has been predicted to have a - -l [33]. but it will be much more complex when other vibrational modes of the adsorbate are involved in the dephasing. Inhomogeneous broadening due to defects and impurities should be characterized by a - -0.5. since the chemical effect on the force constant Should be mass-independent, while the frequency is proportional to mgqu. Tobin [5] argued that the same dependence should occur for domain boundary inhomogeneity. if the lateral interactions are predominantly chemical in nature or dipolar coupling. Therefore the isotopic substitution 14 experiment can be very effective.to.sort out the inhomogeneous broadening. in particular. Third. measurements of the temperature dependence of the linewidth are most widely used and fruitful for determining the line broadening mechanism [5.8.24]. Lifetime broadening by electron-hole pair generation should be essentially independent of temperature. since the characteristic electronic energies for an adsorbate on a metal surface are the Fermi energy of the substrate metal and the width of the adsorbate resonance; the adsorbate resonance level is a few eV wide and is located at a few eV from the Fermi level. The adsorbate resonance is an electronic level induced by the adsorbate near the Fermi level. As the adsorbate vibrates. the resonance is perturbed by the change in the atomic positions of the adsorbate. and its electronic occupation oscillates at the vibrational frequency. resulting in a dynamic transfer of charge between the adsorbate and the metal substrate. In the case of dephasing by interaction with electron-hole pairs. the linewidth should depend on temperature since electrons with energy within kT of the Fermi level can participate. Morawitz [35] has predicted a T‘ dependence for this process. but Volokitin disagrees [32]: he has argued that in. many cases the temperature dependence is negligible. The processes of' phonon. broadening and dephasing are expected to be strongly temperature-dependent in the 100- 500 K range. since the characteristic energies are adsorbate or substrate vibrational energies. The temperature dependence for the phonon mechanism has been theoretically predicted by several people [21.22.32.33]. It is often assumed that inhomogeneous broadening is nearly temperature- independent. since it does not involve dynamics. Its temperature 15 dependence could be uncertain. however. since the nature of the sources of inhomogeneity can vary with temperature. 1.2.2 Summary of previous experiments Vibrations at frequencies above 1000 cm'1 (high-frequency modes) have been most widely studied. mainly because they are relatively easy to measure by IRAS. They include most intramolecular vibrations. particularly the C-0 stretch vibration which is at ~ 2000 cm'l. Even though these modes reveal lots of dynamical and structural information. they do not participate directly in surface processes such as sticking, desorption, and diffusion. Indeed the adsorbate-substrate modes, which are typically at low frequencies (300 - 1000 cm'1 ; called low-frequency modes). are important for these processes. Primarily because these modes are difficult to measure by infrared spectroscopy. there have been very limited studies of linewidths and line shapes of these low frequency modes [17-20,24,38.39]. The lifetime of high-frequency vibrations appears to be determined primarily by the rate of generation of electron-hole pairs in the metal. For CO on Cu(100). Ryberg [40] measured the C-0 stretch linewith as a function of temperature and found that the linewidth and the line shape are temperature-independent, indicating that pure dephasing is not an important line broadening mechanism. He argued that the vibrational lifetime determines the linewidth. and that the primary mechanism of vibrational energy relaxation is the generation of electron-hole pairs in the substrate. The lifetime of the mode is estimated to be T1 =3 1.2 ps. using the measured linewidth of 4.5 cm'1 and the equation T1 - [21rc(Au)]‘1, where c is the speed of light and AV is the full width at half maximum 16 (FWHM) of the line in cm“. Recently. Morin et a1. [14] have directly measured a vibrational lifetime of 2.0 ps for the C-0 stretch mode of CO on Cu(100) using SFG spectroscopy. which would give a homogeneous linewidth of 2.7 cm”. They interpret the energy relaxation process as caused by non-adiabatically generating electron-hole pairs in the copper substrate. The non-adiabatic energy transfer lifetime of this mode has been calculated by a density- functional method [41] and by extrapolation of ab initio Hartree-Fock electronic structure calculations for CO on Cu clusters [42]. Both calculations give lifetimes in the l - 3 ps range, in good agreement with the measured values. Another candidate for lifetime broadening is the l-i-substrate wagging mode of H on W(100) and H on Mo(100) [4] . The first. overtone of the mode has the line shape predicted by Langreth [30] for the electron-hole pair mechanism and its linewidth is temperature-independent. For CO on Ni(lll). CO on Ru(100). and H on Si(100). the observed linewidths of the C—0 and Si-H stretch modes have been attributed to dephasing [4.5]. primarily on the basis of the strong temperature— dependence of the linewidth. In the case of CO, the measured data have been satisfactorily fit to a phenomenological model with several adjustable parameters. For H on Si(100). a first-principles calculation was used to obtain the linewidth within a factor of 2. For low-frequency modes (lower than ~ 1000 cm'l). phonon-mediated decay is more likely to be an important broadening mechanism as discussed in section 2.1.1. However. there are no conclusive experimental measurements so far to show the linewidth caused by the phonon mechanism, l7 and recently Persson and Ryberg [25] have pointed out that all the previous calculations have missed an important factor and the phonon process is less likely to be an important line-broadening mechanism. The first infrared measurement of a low-frequency adsorbate-substrate mode was performed by infrared emission spectroscopy for CO on Ni(lOO) at 475 cm'1 [24] . The measured linewidth was 15 cm'1 at a temperature of 310 K. Unfortunately. the temperature-dependence could not be measured because of the low oscillator strength of the mode and the very limited temperature range accessible with emission spectroscopy, and the line broadening mechanism was not clearly identified. Ariyasu et a1. [21] . and Volokitin et a1. [22] . have calculated the linewidth contribution due to phonon emission decay process for the C-Ni mode of CO on Ni(100) . Ariyasu et al. have used only the two phonon process. while Volokitin et al. have incorporated all phonon processes and found the three-phonon process dominates. They have predicted the linewidth and slightly different but strong temperature-dependences of the linewidth. But Persson [25] has recently pointed out a serious error in all these previous calculations. and has ruled out the possibility of the phonon emission decay for a line- broadening mechanism. The linewidth of the adsorbate-substrate mode of CO on Pt(111). at 467 cm'l, was measured at several temperatures by the same emission spectroscopy [38]. This mode has an oscillator strength 5-10 times greater than that of the corresponding mode of CO on Ni(100) . The results were inconclusive. however. due to non-reproducibility of the measurements and to the limited temperature range available. Recently. the adsorbate-substrate mode of CO on Pt(111) has been 18 measured by Hoge et a1. [20] . Malik and Trenary [l9] . and Ryberg [18,25] . Ryberg has measured the narrowest linewidth. 2.5 cm‘1 at 100 K, compared to 3 - 5 cm'1 at 100 R in the other experiments. suggesting that the other two measurements, as well as the earlier emission measurement, were dominated by inhomogeneous broadening. They also show different temperature dependences of the linewidth; Hoge et al. reported almost no temperature dependence. Malik and Trenary found a strong temperature dependence. from 3 cm‘1 at 80 K to 8 cm”1 at 300 K. and Ryberg found'a linear temperature dependence from 2.5 cm‘1 at 100 K to 4.1 cm'1 at 250 K. Persson and Ryberg [25] have argued that the line broadening is caused mainly by dephasing via anharmonic coupling to the low- frequency parallel frustrated translation and that the vibrational energy decays through excitation of electron-hole pairs. There have also been a few infrared measurements of the low-frequency mode of adsorbed atomic oxygen: an emission measurement [38] of O on Pt(111). at 480 cm'l. a diode laser reflection-absorption measurement [39] .of O on Al(lll). at 580 cm“. and an FT-IRAS measurement [43] of O on polycrystalline Ag, at 351 cm'l. In each case the line is very broad. =38 cm'1 on Pt(111). £100 cm'1 on Al(lll). and =33 cm'1 on polycrystalline Ag. and very weak. Little can be concluded regarding the line broadening mechanism . 1.3 0-0 stretch mode of 02 on Pt(111) In each of the experiments discussed above. there remain ambiguities in determining the line broadening mechanism even for the high-frequency mode. The difficulty arises mostly from the fact that inhomogeneous broadening cannot be ruled out conclusively. Probably. a measurement of 19 the isotopic dependence of the linewidth is the best way to sort out the inhomogeneous broadening, since the mass dependence of the inhomogeneously broadened linewidth is predicted as a - - 0.5 [5] while other line- broadening mechanisms are expected to exhibit a stronger mass dependence [32.33.37]. A measurement of the isotope effect on the linewidth usually requires very high signal-to-noise ratio of the instrument. since the change of linewidth is often very small; the inhomogeneous linewidths for 12C160 and for 12C130 on Pt(111). for example. would differ only by 0.2 cm”. In fact. there have been few successful measurements of the isotope effect on the linewidth. Persson and Ryberg [37] once claimed that the isotope effect of the C-H stretch modes of any) on Cu(100) was measured to be a - - 1. which was taken as an evidence for electron-hole pair broadening. Later these modes were remeasured by the same authors [44] and the linewidths were found to be much narrower than in the original measurement. with little difference in linewidth between the C-H and C-D modes. The difference in linewidth would be 4.3 cm'1 if the line is broadened by electron-hole pair damping. or 1.8 cm'1 if the inhomogeneous broadening is dominant. ‘The only clear experiments have been for adsorbed hydrogen. for which the mass and frequency change are so large that it cannot be assumed that the chemical interactions and decay process are the same for both isotopes. We have chosen 0; on Pt(111) as a first system to study the line broadening mechanism. Here. we have to admit that this is not the kind of experiment the instrument was designed for --- it is not a very low frequency mode and it is not an adsorbate-substrate mode either. But. oxygen on Pt(111) is important. technologically for heterogeneous 20 catalysis. such as the oxidation of CO. NO and NH3. as well as scientifically for the study of dissociative chemisorption. The 0-0 stretch band is also a little bit easier to measure; the oscillator strength is larger than that of most adsorbate-substrate modes. but much smaller than C-O stretch. and it is at moderate frequency, ~ 875 cmTK Oxygen on Pt(111) has been widely studied by many surface experimental techniques; temperature-programmed desorption (TPD) [45-49]. low energy electron diffraction (LEED) [47]. X-ray photoelectron spectroscopy (XPS) [50], near-edge X-ray-absorption fine-structure spectroscopy (NEXAFS') [51]. and molecular beam scattering [52]. The system has been relatively less studied by surface vibrational spectrosCOPY: only one reflectionrabsorption-infrared spectroscopy measurement [53] and two electron energy loss spectroscopy measurements [46,47] have been reported. At low temperature (5 ~100 K). oxygen adsorbs molecularLy on the Pt(lll) surface, with its bond axis parallel to the surface [46,47]. This molecular oxygen is a metastable precursor state to dissociation into atomic oxygen. which forms an ordered p(2x2) overlayer upon heating above 150 K. Despite the orientation of the 0-0 bond. the O-O stretch vibration is observed at 875 cm'1 with a relatively large dynamic dipole, indicating that large adsorbate-substrate charge transfers occur as the bond stretches. The 0-0 stretch frequency of the chemisorbed Oz. 875 cm'l, is surprisingly lower than the corresponding frequency of gas phase 02, 1556 cm’l. In fact. the double bond of free 02 is reduced to a single bond upon chemisorption on Pt(111) surface. This reduction is explained by a large static charge transfer from the metal substrate to the antibonding 21r' 21 orbital of the adsorbed 02 molecule. possibly changing the orbital from the half filled state to the filled state [45.47]. Canning and Chesters [53] have observed the 0-0 stretch band of molecularly adsorbed 02 on a recrystallized Pt foil at 875 cmfl, with a relatively large linewidth 20 - 25 cm'1 at the saturation coverage 0 - 0.4. They have measured the frequency shift due to an isotopic mixture and confirmed that the band is really due to molecularly adsorbed oxygen on the surface. They have also observed the lowest frequency 845 cm’1 at a _coverage 0.2. and have explained that the overall frequency shift of 30 cm’1 is caused by dipole-dipole coupling. Persson [54] has argued that the line broadening of the O-O stretch vibration is caused by the generation of electron-hole pairs in the substrate, based on an analysis of the line shape using Langreth's model. The possibility of inhomogeneous broadening is excluded. by showing that the same line width has been deduced from EELS and IRAS data. but the experimental evidence is indirect and.inconclusive. Furthermore. the IRAS spectra were measured on a recrystallized foil. without surface analytical capabilities. The dominant line-broadening mechanism can be determined experimentally by measuring the temperature-dependence and/or the mass dependence of the linewidth, and by examining the line shape. As discussed in 1.2.1. the line will be either highly asymmetric or a symmetric Gaussian.if it is inhomogeneously'broadened” while the line will be asymmetric with a specific lineshape and an extended tail on either side if electron-hole pair damping dominates the broadening [30]. Unfortunately. we cannot investigate the temperature-dependence of the 22 linewidth of the 0-0 stretch mode. at 875 cm”. since the desorption of 02 molecules already starts to occur at 100 K. The isotopic dependence of the linewidth is, however. expected to be relatively large; the linewidths for 1602 and for 1302 would differ by 1.2 cm'1 if inhomogeneous broadening is dominant. or by 2.5 cm'1 if electron-hole pair damping dominates the linewidth. The difference of 1.2 cm'1 or 2.5 cm'1 can be measured by IRAS. if the system has high noise-to-signal ratio and excellent baseline substraction. and the measurement will shed light on further understanding of the line broadening mechanism. The measurement of the isotopic dependence of the linewidth for 02 on Pt(111) has been thus chosen as the first project of our new instrument. The Pt-O vibration of the atomic oxygen on Pt(111) has been measured at 480 cm'1 by EELS [46.47] and IR emission spectroscopy [38]. For molecularly adsorbed 02 on Pt(111). the Pt-Oz vibration has been measured at 380 cm'1 by EELS [46.47]. but it has never been measured by infrared spectroscopy. Because of the low temperature required, emission spectroscopy could not be used for the Pt-Oz band. These bands are expected to be very weak and difficult to observe with conventional IRAS instrument. With a new apparatus spanning the frequency range 350 to 3500 cm'l. these modes will be observed so that the entire process of adsorption and desorption of such molecules as 02 can be traced with high resolution. 1.4 Summary In this thesis work. a new far-infrared reflection-absorption spectroscopy (IRAS) system has been built using a grating spectrometer and a thermal source. The system has been specifically designed for maximum sensitivity in the 330 - 1000 cm'1 spectral range (though it is usable up 23 to >3500 cm'l) with better than 2 cm'1 resolution. It incorporates a liquid-nitrogen-cooled grating spectrometer and other cooled optics to minimize ambient blackbody radiation. The spectroscopic system is coupled to an ultrahigh vacuum surface analysis chamber. Most of the thesis is about the design. construction and.evaluation of the IRAS instrument, with the preliminary measurements using the system. The rest of the thesis is organized as follows. In chapter 2. the feasibility and design criteria of the IRAS system are presented. The typical IRAS signal on a platinum surface and background photon noise are calculated and analyzed. Characteristics of some optical elements and techniques are discussed with a viewpoint of maximizing the signal-to-noise ratio. The design of our IRAS system is discussed. specifically for operation at frequencies less than 1000 cmfih In.chapter 3. I describe the details of the apparatus and how’we have built the system. In chapter 4. the performance of the system is evaluated and compared with the design goal and other systems. along with the calibrations of the detector and spectrometer. and with the 'quantitative analysis and measurements of the system noise. The preliminary measurements of CO on Pt(111) and.02 on Pt(111) are presented in chapter 5 and conclusions follow in chapter 6. 24 References 10. 11. 12. 13. 14. G.A. Somorjai. Chemistry in Two Dimension: Surfaces. (Cornell University Press. Ithaca. New York. 1981). A. Zangwill. Physic5' at Surfaces. (Cambridge 'University' Press, Cambridge. 1988). 6.8. Higashi. Y.J. Chabal. G.W. Trucks and K. Raghavachari, Appl. Phys. Lett. 56. 656 (1990). Y.J. Chabal. Surf. Sci. Reports 8. 211 (1988). R.G. Tobin, Surf. Sci. 183. 226 (1987). A.M. Bradshaw and E. Schweizer. in .Advances in Spectroscopy: Spectroscopy of Surfaces, edited by P.E. Hester (Wiley. New York. 1988) pp. 413 - 483. F.M. Hoffmann. Surf. Sci. Reports 3, 107 (1983). J.W. Gadzuk and A.C. Luntz. Surf. Sci. 144. 429 (1984). I. Ueba. Progr. Surf. Sci. 22. 181 (1986). E.J. Heilweil. M.P. Casassa. R.R. Cavanagh. and J.C. Stephenson, J. Chem. Phys. 82. 5216 (1985). J.D. Beckerle. R.R Cavanagh. M.P. Casassa. E.J. Heilweil, and J.C. Stephenson. J. Chem. Phys. 95. 5403 (1992). R.R. Cavanagh. J.D. Beckerle. M.P. Casassa. E.J. Heilweil. and J.C. Stephenson. Surf. Sci. 269/270. 113 (1992). A.L. Harris, L. Rothberg, L.H. Dubois. N.J. Levinos. and L. Dhar. Phys. Rev. Lett. 64, 2086 (1990). M. Morin. N.J. Ievinos. and A.L. Harris, J. Chem. Phys. 96. 3950 (1992). 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 25 P. Guyot-Sionnest. P. Dumas, Y.J. Chabal and G.S. Higashi, Phys. Rev. Lett. 64. 2156 (1990). P.L. Richards and.R.G. Tobin. in.Vibrationa1 Spectroscopy of Molecules on Surfaces. edited. by .J.T. ‘Yates. Jr. and. T.E. Madey (Plenum Publishing Co. New York, 1987). C. Hirschmugl. G.P. Williams, F.M. Hoffmann and Y.J. Chabal, Phys. Rev. Lett. 65. 480 (1990). R. Ryberg, Phys. Rev. B40. 5849 (1989); B40. 8567 (1989). I.J. Malik and M. Trenary. Surf. Sci. 214, L237 (1989). D. Hoge. M. Tushaus, E. Schweizer and.A.M. Bradshaw. Chem. Phys. Lett. 151. 230 (1988) J.C. Ariyasu. D.L. Mills. R.G. Lloyd. and J.C. Hemminger, Phys. Rev. 328. 6123 (1983); Phys. Rev. B30. 507 (1984). A.I. Volokitin. O.M. Braun and V.M. Yakovlev, Surf. Sci. 172. 31 (1986). V.P. Zhdannov. Surf. Sci. 197. 35 (1988). S. Chiang. R.G. Tobin. P.L. Richards and P.A. Thiel. Phys. Rev. Lett. 52. 648 (1984). B.N.J. Persson and R. Ryberg. Phys. Rev. B40, 10273 (1989). P. Dumas. Y.J. Chabal, and G.S. Higashi, Phys. Rev. Lett. 65. 1124 (1991). R. Disselkamp. H.-C. Chang and.G.E. Ewing. Surf. Sci. 240. 193 (1990); C. Node. R.R. Richardson and G.E. Ewing. J. Chem. Phys. 92, 2099 (1990). B.N.J. Persson. Solid State Commun. 27. 417 (1978). 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 26 B.N.J. Persson and M. Persson, Solid State Commun. 36, 175 (1980); Surf. Sci. 97, 609 (1980). D.C. Langreth, Phys. Rev. Lett. 54. 126 (1985). Z. Crljen and D.C. Langreth. Phys. Rev. B35. 4224. (1987). A.I. Volokitin. Surf. Sci. 224. 359 (1989). B.N.J. Persson. J. Phys. 017. 4741 (1984). B.N.J. Persson and R. Ryberg, Phys. Rev. B32. 3586 (1985). H. Morawitz, Phys. Rev. Lett. 58. 2778 (1987). P. Jacob. Y.J. Chabal and K. Raghavachari, Chem. Phys. Lett. 187. 325 (1991). B.N.J. Persson and R. Ryberg. Phys. Rev. Lett. 48, 549 (1982). R.G. Tobin and P.L. Richards. Surf. Sci. 179. 387 (1987). V. M. Bermudez. R. L. Rubinovitz and J. E. Butler, J. Vac. Sci. Technol. A6, 717 (1989). R. Ryberg, Phys. Rev. B32, 2671 (1985). T.T. Rantala and A. Rosen. Phys. Rev. BB4. 837 (1986). M. Head-Gordon. J.C. Tully, J. Chem. Phys. 96. 3939 (1992). X.-D. Wang. W.T. Tysoe, R.G. Greenler. and.K. Truszkowska. Surf. Sci. 257. 335 (1991). R. Ryberg. J. Chem. Phys. 82. 567 (1985). J.L. Gland. Surf. Sci. 93. 487 (1980). J.L Gland. B.A. Sexton and G.B. Fisher. Surf. Sci. 95, 587 (1980). H. Steininger. S. Lehwald and H. Ibach. Surf. Sci. 123. l (1982). A. Winkler. X. Guo. R.R. Siddiqui. P.L Hagans and J.T. Yates. Jr.. Surf. Sci. 201. 419 (1988). C.T. Rettner and C.B. Mullins. J. Chem. Phys. 94. 1626 (1991). 50. 51. 52. 53. 54. 27 A.C. Luntz, J. Grimblot and D.E. Fowler, Phys. Rev. BB9, 12903 (1989). W. Wurth, J. Stdhr. P. Feulner. X. Pan. K.R. Bauchspiess. Y. Baba. E. Hudel, G. Rocker. and D. Menzel, Phys. Rev. Lett. 65, 2426 (1990). A.E. Wiskerke, F.H. Geuzebroek. A.W. Kleyn and B.E. Hayden. Surf. Sci. 272. 256 (1992). N.D.S. Canning and M.A. Chesters. J. Elect. Spect. Relat. Phenom. 29, 69 (1983). B.N.J. Persson. Chem. Phys. Lett. 139, 457 (1987). 2. Feasibility of the IRAS Experiment In this chapter. I discuss the characteristics, feasibility and designs of infrared spectroscopic systems that are required for the studies of adsorbate vibrations at frequencies below 800 cmfih Quantitative analyses are performed. to show that such a system is feasible. using a conventional thermal infrared source. together with a modern infrared detector. a cooled spectrometer and cooled optics. A. specific design of such an infrared reflection-absorption spectroscopic system is presented. 2.1 Comparison of IRAS with other vibrational spectroscopies Surface vibrational spectroscopy has been one of the most important techniques to study surface dynamics and adsorbates on solid surfaces. Many different kinds of surface vibrational spectroscopy have been developed for this purpose [1]. For the present discussion. I compare three techniques below. that is. electron energy loss spectroscopy [2], infrared emission spectroscopy [4.5]. and infrared reflection-absorption spectroscopy [1.6.8.9]. Electron energy loss spectroscopy (EELS) uses the inelastic scattering of an electron beam. A monochromatic low energy electron beam (-5 eV) is incident on the surface and the scattered electrons are detected with an energy-selective detector so that the energy loss due to adsorbates is measured. EELS is one of the most widely used surface vibrational spectroscopic tools because of high sensitivity as well as a wide spectral range (200 - 4000 cm“). Also it can detect modes that are not infrared-active. Virtually all adsorbate vibrations of interest 28 29 except the ones below 200 cm"1 can be and have been measured by this technique. The resolution, however, is typically 40 - 80 cm'1 (resolution of 8 - 16 cm“1 has been achieved recently [3]) which is not good enough to measure the linewidths and frequency shifts, which are often <10 cm'l. and of which accurate measurements are very important for the study of surface dynamics. Another drawback of EELS is that it is very difficult to evaluate intensities quantitatively, because the scattering processes are not well understood. Infrared emission spectroscopy [4,5] routinely achieves a resolution of l - 5 cm'l. and covers much of the relevant frequency range. from 350 to >2500 cm'l. with good sensitivity. In the emission technique. the thermal equilibrium radiation emitted by the sample surface itself is collected. dispersed by a liquid-helium-cooled monochromator. and detected with a sensitive photoconductive detector. Comparing a spectrum measured with a clean surface with one measured with an adsorbed overlayer, the small enhancement of emissivity due to the adsorbate vibration can be detected. Since the sample itself is used as the radiation source. the sample temperature needs to be high enough to maintain adequate sensitivity. It has to be higher than 200 K even at 500 cm“. At higher frequencies. above the peak of the blackbody curve. the emitted intensity falls exponentially with decreasing temperature. and even higher temperatures are required to get the necessary sensitivity. Since many adsorbates begin to desorb at temperatures below 400 K, and often below 300 K. the temperature range available to experiment is very narrow; Even for such a relatively stable system as CO on Pt(111) the usable 30 temperature range for the emission technique extends only from 200 to 275 K. Infrared reflection-absorption spectroscopy (IRAS) [1.6.8.9] uses a separate infrared radiation source and thus does not impose any constraints on the sample temperature. Infrared radiation from the source is reflected off the sample, sent to a dispersive monochromator or Fourier transform infrared (FTIR) instrument and collected by a detector. The spectrometer can come before the sample. The adsorbed layer on the surface absorbs a small amount of radiation so that the reflectance of the surface is changed. The small change in the reflectance is detected by comparing a spectrum measured with a clean surface with one measured with an adsorbed layer. IRAS routinely achieves resolution of 1 - 5 cm'l. Conventional IRAS. which uses a room temperature spectrometer and optics. however, is difficult to extend to frequencies less than 800 cm'l. We have built a new system which maintains the sensitivity in the low frequency region by using a sensitive detector and reducing the background photon noise from the environment by cooling the spectrometer and the relevant optics. This issue is discussed in detail in sections 2.5 - 2.7. 2.2 Surface selection rule on metal surfaces In an IRAS experiment. the electromagnetic field of the infrared radiation interacts with the dynamic dipole moment associated with a particular normal vibrational mode of the molecules and a small proportion of the radiation is absorbed by the vibrational excitation of the molecules. Since the dielectric response of the substrate dominates the electromagnetic field on a surface at infrared frequencies. it is important to investigate the dielectric properties of substrates to 31 optimize the experimental conditions. On metal surfaces. the s-polarized component of incident infrared radiation is almost perfectly screened and only the p~polarized component survives and is enhanced by a factor of nearly 2 for angles of near- grazing incidence. as was first recognized by Francis and Ellison [7]. As shown in Figure 2.2.1 the s- and p-polarized electric fields of the radiation are perpendicular and parallel to the plane of incidence. respectively. Figure 2.2.2 shows the dependence of the electric field strength E/Eo on the angle of incidence 0 for both s- and p-polarized light at a platinum surface for u - 500 and 2100 cm'1 [8]. Here. E0 is the amplitude of the incident beam and E the resultant amplitude of the electric field at the surface. The p-polarized component B, can be further split into components EIn and EPI’ perpendicular and parallel to the surface. respectively. The field strength of Em increases with increasing angle of incidence 0, reaching a maximum between 80 and 90° but falling rapidLy to zero at 90°. On the other hand s-polarized light and the .parallel component of p-polarized light are screened.out to be negligible, particularly at 500 cm“, as is also shown in Figure 2.2.2. Since the parallel component of the field is screened at the metal surface. any molecular dynamic dipole moment parallel to the metal surface is screened out. This is visualized by an induced image dipole in Figure 2.2.3. For the dipole perpendicular to the surface the apparent dipole moment is enhanced since the image dipole is in the same direction. while the parallel dipole moment is cancelled out. It is. therefore. expected that only vibrational modes with dynamic dipole perpendicular to the surface should be observed in IRAS. 32 Recently. however, a frustrated rotation mode [10] has been observed with IRAS. Persson and Volokitin [11] have theoretically shown that parallel translation modes can be observed in IRAS, as anti-absorption peaks or strong Fano lineshapes. due to indirect excitations of electrons in the metal substrate. 2.3 Incidence angle dependence of the surface signal Most IRAS analysis is based on the three-layer model introduced by Greenler [12], who investigated the conditions necessary to give the highest absorption in a multiple reflection technique. Later. however, it was shown that a single reflection was sufficient to obtain good infrared reflection spectra [13]. Since then the single reflection technique has been extensively used and refined for measuring IRAS of adsorbates on metal surfaces. The macroscopic theory of IRAS has been refined by McIntyre and Aspnes [14], Dignam et a1 [15]. Bagchi et a1 [16]. Persson [17] . and Langreth [18] . Recently Tobin [19] used a macroscopic theory of reflectivity developed by Langreth [18] to incorporate the effect of the dielectric response of the substrate into the IRAS analysis. For present purposes it does not matter which analysis is used to calculate the adsorbate signal. since we are not trying to do any detailed quantitative work. We follow the derivation by Persson. Classical electromagnetism is used to calculate the reflectance of infrared radiation on a metal surface. The macroscopic response of a solid surface is described by a dielectric function 600) which is a prOportionality factor between the electric displacement D and an applied electric field E. In a linear medium the relation between the two is: D- 6E . (2.3.1) 33 First we consider a two phase system with a plane boundary of vacuum and bare substrate. The infrared beam is described by a plane electromagnetic wave through a medium. Suppose p-polarized light is incident at an angle 0 on a surface from the vacuum side as shown in Figure 2.3.1. The incident light is described by an electric field: E. - Eo (xcoso + fisinO) ei‘k't’m’ . (2.3.2) The reflected field is given by : Er - Eo rp (-§icoso + fisinO) ei‘k'T'm" , (2.3.3) where r is given by the Fresnel equation for p-polarized light [1]: P ecoso - (6 - sin20)1/2 r - , (2.3.4) ecoso + (e - sin20)1/2 So the reflectance is given by: 11,- lrplz. (2.3.5) The intensity of the reflected wave is given by: c c Er'E;I- 8x 8x |::,,|21~:.,2 . (2.3.6) The field in the z-direction, which is perpendicular to the surface, at the surface (z-O) is: E“ - E0 sin0 (1+rp) e1‘kx'mo’“) . (2.3.7) Now suppose the surface is covered by a very thin (d<qoomocmuuam :o_LQcomuo _osco;_ Luggage vac cm~mco_o¢ 388:8QO E58; $9: .mEoEm 838-5 3: ; co_Lucha_o cocmum_w >Ocmcm gem Loamo>co moczom m_ U885 82% case \lllllumar comummoa umc0c_:_ m”_m wm_ooulmzm seem 538523 9:95 Eosasoszaa: 3339 :2. 838.23 S u_qcem _oum>cu m_o:_m ;L_a cwpsczu m_m>_o:< ouoaczm ssaoo> ;o_;oLL_: x as - a as; Beam :8 m; ”Ezeamm TE 83 - 8m ”38.. 858% 555 EoumoEBam “Giza Swizz -mfi 98 Liquid nitrogen Electrical / feedthrough / feedthrough .v :L, LT “E III Eli—:3 t. -L if W ““1 _ Liquid /_ nitrogen can / F \— Cold plate . tical /— Supporting rod .==;: /— a?“ Kinematic mount = E 117—— Teflon base ....................... ........ ............... ....................... ....................... ....................... ....................... ....................... II‘IIIILIIAAJAJAIILILJL to ion pump Figure 3.1.1 Schematic of the source cryostat mmbmoxuo oOLSOm 23 tom mot/.50 wczooo N. _ .m “:st A2505 o5? . v m m a o ) in"?! — q q — d — d — u > I n .v I l .. .e b s :1 etc .‘I 0'. (9 file) i ’ titli“ -l . . '3 maize: moaDOm I Eme H5339: I .5925 I 3505 cameos? III Baa omen. I fifim fixo I OOH com com ()1) arnieredmai 100 .\ Flansed teflon bearins .................. .................. .................. ............ /— Setscrew /—— Globar 3? /— Macor insulator 3?? m Cold plate T v l lllllll \ . ‘ .'. Figure 3.1.3 Schematic of the source housing 101 83% gen 05 we «woke..— véd arm someone gen newsflash hone—u 48% ”355 32a 28 _ 0 g a a . .wv -. ................................. m ..... I ........... 338 .558 B.E.—E 133350 deuce—«2 aces— new mesh:— 3 03968 Goran—:5 .83: EH0 new flea II\. 102 cold copper legs -- connected to \per feedthroughs . . 5w“ .......................... ....................... ......................... 42:25:52: thermocouple ;;;;;;;;;:;;; Mam” Pi°°° ............. ............ .............. .......................... ............ .............. um amp ............ . ........ . . . . . .,..... ..._.:.:.:.:. .'.'.'.'.'.'.,.,.... 0.. ....... . . . . . . ..... .‘9.0.. . . . L; . £ . . . . . . . .Q.Q.g.. I": I: .. .. .. ' Z l-i‘ “2'! ’ '- .. . a: .'l' 3.1. . ... 15‘ z A. - . . . I. I I'I .CIDI‘ TOP VIEW ....... ............ ....... ...... ....... ...... ...... ...... sample sapphire washer I ? Wm mm mm tungsten filament heater Figure 3.3.1 Schematic of the lower part of the sample mount. 103 threaded ring f O ' ring holder atmosphere side / V J g .-:?:1;3:1:3:3:3:i:4;i.1:3:3.?:i:1:3:3;1:3:3:':322323:?:313:3:3:3:?:3:3:3:3:?:13:1: // ¢ efsiz§é§2§3§5§3§3§2§2§3§5§2§2§;§if2§2§:§2§s§s§2§2§3§3§3§2§5§3§3§2§5§3§3§3§3§3§3§2§s§:3;/ / ............................................ a + ”M // UHV‘t side retaining ring differential pumpins port 4. 5" Conflat flange CsI window Figure 3.3.2 Differentially pumped CsI window. Vacuum seals are provided by three viton O-rings. The CsI window is pressed on only by the O-rings Nomalized flux 104 0.5 - 0.3 - 0.2 r 0.0 . 1 1 1 —1.5 —0.5 0.5 Distance (cm) Figure 3.3.3 The calculated uniformity of the gas doser. 1.5 105 non—caches? on..— .«e deacon each. in?!» aid 953m \\ omen deuce ages" emu—“Bonu— =enn 8389333 \ \ henchmen \ 303a defiance." men 5233 .8an ope—Q 93 393.98 3.3a omen seven—38% «aid 33.. 38 388 use couch—nu Boo: ...................................... ...u........na.”"umunwmnmuummuw.mn.unmwuumumenswewumwww.n.UEwemomma?“”wan“.u “nun“. i... M 33 5.0 Bebe «fleas—u 106 spectrometer base plate cryostat top plate angled bracket 1/4" pipe thread quick disconnect clampscrewtipped withballbearing brazed radiation shield spectrometershell Figure3.4.2Diagramoftheclampmechsnism. Threescrewsareusedto clamp the spectrometer base plate to enhance the mechanical stability. 107 cm LeboEobocmm on: com mo>czo mczooo med madman $.5on one; ON 0“ . . mocha mLmEmo EoEm :oScLUeL mama omen, Eon? imm mcficcm cocci: wcwEoH on: com ()1) ainqeiedmei 108 coax—6:89 2: no 26?. .828 v.96 2:3..— (3 «E —‘B. Sui-he‘s. hurt; I 109 spectrometer base plate . Figure 3.4.5 Grating table assembly with Inductosyn rotary position transducer. ' 110 Figure 3.4.6 Schematic of the grating drive 111 LN2 dewar edge-welded bellows Quick clamp copper block ellipsoidal mirror mirror mount adjusting knob screw tipped with spring loaded ball bearing vacuum can radiation shield Figure 3.5.1 Schematic of the liquid-nitrogen-cooled transfer optics. 112 ...................... RELAY afsmscr # 100 I :~J\/W\/\/\——+15 E g s i I? : I . “WW—\i 3 1'1.— _/._/WW_.-E e INPUTTEST - -.-._R.f-__.-._. " DETECTOR -c-u--------------- .......................................................................................... Figure 3.6.1 The circuit diagram of the Si:B detector amplifier. 89593 593 £3 3888 as .838. genes: as. to sauce seem Se team m3. anon 3:888 .838 3.33- 113 .8938 8.3.83 moan an 8:882 gov e3. .335 couscous hoe—econ .80st £338 .888 mafia—3n 83089539 4. System Performance We have built our instrument specifically for measuring low frequency surface vibrational modes at submonolayer coverage, which generally have much weaker signals, often by orders of magnitude, than high frequency modes. The central issue is therefore the capability to detect a small signal on a large background, which requires a sensitivity better than ~1 x 10" in the frequency region below 1000 cm“. In fact, better sensitivity (~1 x 10-5) is achieved with FTIRs in the high frequency region (~2000cmf1), which is possible because the detectors are not sensitive to low frequency radiation, so they are not subject to background photon noise. This chapter presents a quantitative analysis of the performance and limitations of our system, along with comparisons with other systems. 4.1 Calibration of spectral frequency Accurate frequency measurements are important in IRAS to locate the absorption band accurately. The calibration of frequency can be performed usually by two methods; well-known absorption spectra are measured and compared with their standard spectra, or spectra of different orders for a known monochromatic source are measured and compared with calculated ones. We have used the latter method to calibrate the spectral frequency. He-Ne laser spectra of several orders are measured and used for the calibration of the spectral frequency. We place the laser in front of the window in the source cryostat and adjust the mirrors and sample as we do for the system alignment. An empty slit is selected at the filter wheel to allow the laser beam to pass through the system. As the grating is rotated from 0° to high angles, sharp laser spectra of many orders are 114 115 measured at the detector. The diffracted angle of the He-Ne laser line of n-th order is expressed as the grating equation: nA - 2dsin0ncosd, (4.1.1) where A is the wavelength of the laser beam; 6328 A, n the order of diffraction, d the groove spacing of the grating, 0n the grating angle at the n-th order peak, and ¢ the half angle between the incident and diffracted beams of light at the grating, as shown in Figure 4.1.1. We carefully measure laser spectra of several orders along with on, and plot nA versus 2dsin0m to get cosd from the slope, using least squares fitting, as shown in Figure 4.1.2. In this way, we have determined cos¢ both for the cold and the room temperature spectrometer; cos¢ is 0.982 1 0.003 for the cold spectrometer and 0.986 :t 0.003 for the room temperature spectrometer respectively. The expected value of cos¢ is 0.9869 from the design value of ¢ - 9.302° at room temperature. The uncertainties are estimated from the root sum square of the products of the standard deviation of each data point multiplied by the effect which the data point {2dsin0n} has on the determination of the parameter cosd, assuming that the uncertainties are instrumental and that the standard deviations for the data points are equal [1] . The apparent change of cos¢ with temperature is mainly due to the change of the groove spacing d of the grating. The change of cos¢ due to the change in d is estimated to be 3.85 x 10'3 as the temperature changes from room temperature to 80 K, using the thermal contraction AL/L - 3.9 x 10’3 for the aluminum substrate. This agrees well with our measured change of 4 x 10‘3. 116 4.2 Spectrometer resolution As discussed in 2.4, the angular resolution d0 of the monochromator is determined by the slit width w and the effective focal length f of the camera mirror: d0 - w/(2f), (4.2.1) where a factor of 1/2 comes in because the angle between the incident and diffracted beams changes by an amount 2d0 as the grating angle 0 is changed by an amount d0, for a given wavelength. It is assumed in the equation (4.2.1) that the number of grooves illuminated is much larger than the resolving power, and that aberrations are small compared to the slit width w. The first condition is easily satisfied in our system - the number of grooves illuminated is always >9000, compared to a typical resolving power ~400. The second condition was verified by a ray-tracing analysis, by visual inspection of the image, and by direct resolution measurements. The angular resolution of the spectrometer was determined for each 'slit width by measuring the full width at half maximum for the zeroth order peak, using a He-Ne laser as the light source. The laser beam was expanded to fill the optics. The zeroth order peaks for different slits are shown in Figure 4.2.1 and the angular resolution is plotted as a function of slit width in Figure 4.2.2, along with the calculated values using the equation (4.2.1). The measured resolution behaves as if the effective focal length were 74 cm instead of 65 cm. The reasons for the disagreement are not well understood, but possibly because the slit was not filled completely with the laser beam, or the effective widths of the slits are reduced by ~ 10 1 due to black paint coating. 117 The frequency resolution.af the monochromator du can be obtained from the grating equation, the focal length, and the slit width. Let us start with the grating equation, for a fixed angle 2d between the incident and diffracted beams: nA - 2dsin0cos¢ , (4.2.2) where A is the wavelength of the light, n the order of diffraction, d the groove spacing of the grating, 0 the angle of the grating, and d the half angle between the incident and diffracted beams at the grating (see Figure 4.1.1). Differentiating with respect to 0, we get an equation: n dA - 2dcos€cos¢ d0 . . (4.2.3) Dividing equation (4.2.2) by equation (4.2.3) gives: dA du d0 A v tanfi (4.2.4) Assuming that we use a grating of 300 grooves/mm (60 grooves/mm) the grating angle 0 is 48.S9° (36.87°) for u - 2000 cm‘1 (500 cm“). With a 2 mm wide slit and thus the measured angular resolution 0.075° (see Figure 4.2.2), the frequency resolution.du and.the resolving power R (- u/du) are 2.3 cm'1 and 870 at 2000 cm'1 (du - 0.9 cm’1 and R - 570 at 500 cm’l). 4.3 Detector calibration The detector receives a beam of infrared radiation and converts it into an electrical signal. As mentioned above, we use an extrinsic Si:B photoconductive detector. Photons incident on the detector excite electrons from the valence band to an impurity band, which is formed by the boron impurities, leaving holes in the valence band so that the electrical resistance changes as the number of incident photons changes. Due to this quantum nature, photoconductive detectors are sensitive and 118 respond faster than thermal detectors, but they have a well-defined spectral limit beyond which they will not respond. The long-wavelength cutoff is determined by the semiconductor band gap for the intrinsic case or by the depth of the impurity band for the extrinsic case. The responsivity of a detector is a measure of its sensitivity; the responsivity of a photoconductive detector can be expressed as the ratio of the detector photocurrent to the incident radiation power wattage or the number of incident photons per second. Since the responsivity is different from one detector to another even for the same type of detector, it is important to calibrate the detector in order to understand the sensitivity of a system and the nature of noises. The calibration setup for the Si:B detector is shown in Figure 4.3.1. The blackbody source is home-made of a copper cone (1.5" diameter, 3" long) painted black with a high temperature black paint, and fitted into a ceramic cone surrounded by a Nichrome heater coil. The heater is plugged in a Variac and the temperature is measured with a K-type thermocouple attached to the copper cone. The source aperture is an aluminum sheet with a 0.25" diameter hole (or 0.125" diameter hole). A rotating chopper is placed behind the source aperture to chop the radiation. The output signal voltage is measured with a PAR S209 lock-in amplifier using a reference signal from the chopper controller. The voltage output (peak-to-peak) at the amplifier of the detector can be expressed as: vout ' Mgnfpchop . (4-3-1) where 3 is the current responsivity of the detector (here the detector includes the Winston cone and cavity), As the gain of the 2nd stage 119 amplifier (see section 3.6), Rt the feedback resistance and Pchop the chopped power from the radiation source. Here the detector photocurrent IPflmp due to the radiation power and bias voltage flows only through the feedback load resistance and is measured as a voltage me at the output of the amplifier circuit. In extrinsic photoconductive detectors, a fraction of the incident photons produces either conduction electrons or holes, which generate a photocurrent under the influence of an external bias voltage. The conduction electrons and holes are called charge carriers or carriers and the number of carriers generated.per incident photon is called the quantum efficiency 0. All the carriers generated by incident photons do not necessarily reach the electrode. Some carriers may decay by recombination before reaching the electrode. The ratio of the number of carriers reaching the electrode to those generated by incident photons is called the photoconductive gain G and can be expressed by [2]: G - ___ , (4.3.2) where 1 is the carrier lifetime and t1. the transit time of a carrier from one electrode to the other one. The photocurrent is therefore given by the equation: I - Gnflchope , (4.3.3) where G is the photoconductive gain, 0 the quantum efficiency (0 includes the transmittance of the Winston cone and cavity in our analysis), Nam? the chopped photon rate from the radiation source, and e is the electronic charge. The photo current flows through the feedback load resistor. The output voltage at the amplifier of the detector is therefore given by: 120 v0“, - AstG’nNchope . (4.3.4) While As and R, are known from the amplifier circuit of the detector, the chopped power and photon rate are calculated from the source temperature and the geometry of the calibration setup. Since the radiation source is modulated by a room temperature chopper and therefore the detector sees alternately the radiation source and the room temperature chopper blade, the chopped power (photon rate) is the difference of the power (photon rate) from the blackbody radiation source and that from the chopper blade. The power emitted by a blackbody into an effective solid angle 0,“, in both polarizations, at frequencies between ml and 4:2 is given by the equation (see section 2.5): 2“ (02 U3 P(T) - an,“ —— ___ dw (2K)3c2 wl e “V“ - 1 - A0.“ dx , (4.3.5) 2h (kT)‘ x2 x3 (2103c2 ft‘ Lg e‘ - 1 where A and T are the area and temperature of the blackbody, and x - fiw/kT. The corresponding photon rate emitted from a blackbody is calculated by the equation: 2 (kT)3 [x2 x2 (2«)3cz #3 dx . (4.3.6) N(T) "' An"! x 1 x1 3 ‘ The throughput is calculated from the geometry of the calibration setup (see Figure 4.3.1): A1A2 A0.“-————, (4.3.7) d2 where A1 and A2 are the detector and source aperture areas respectively, and d is the distance between the detector and source apertures. Once Vent 121 is measured at the detector output and Pchop is calculated, the current responsivity of the detector is computed from the equation (4.3.1). The quantity Gr; can be obtained from the equation (4.3.4) if Vout is measured at the detector and Nchop is calculated using the equation (4.3.6). The responsivity a of our detector has been measured for three different source temperatures and the results are shown in Table 4.3.1, along with the quantity Cry. The measured responsivity 3 is 0.14 - 0.15 (A/W) and the measured value of Ga is 0.92 x 10'2 - 1.01 x 10'2. The transmittance of the filter and window is incorporated in calculating the photon rate and radiation power, while the efficiency of the detector optics, Winston cone and cavity, is included in G!) and a as mentioned above. Some parameters of the calibration setup used for measurements are as follows: aperture of the detector (8 mm dia.) : A1 - 0.50 cm2 aperture of the source (1/8" dia.) : A2 - 7.9 x 10'2 cm2 distance between the two apertures : d - 10.9 cm gain of the 2nd stage amplifier : A! - 10 filter frequency : u - 336 ~ 725 cm'1 filter transmittance : t: - 0.72 KRS-S window transmittance : tw - 0.70 chopping frequency : f - 92 Hz detector bias voltage : Vb“, - 6.0 V In the rest of this section, we discuss the noise-to-signal ratio of the calibration setup for the detector, in order to separate the quantum efficiency 1) and the photoconductive gain G from the measured quantity On. The discussion about noise here is solely for the calibration setup of the detector. Noise for the IRAS system will be discussed in sections 4.5 - 4.7. In an arbitrarily given time interval t, the number "fit of carriers are generated in the detector by incident photons. Since the photons 122 arrive at the detector in a random fashion, Poisson statistics can be used to calculate the root-mean-square (RMS) fluctuation in this number of charges, that is, A(nNt)m - (nNt)1/2. The time averaged RMS noise current is e(nN/t)“q. Since the postdetection noise bandwidth is given by l/(2t) for averaging over a time t, the corresponding noise current per unit post detection bandwidth is: AIm - Ge(4nN)1/2 (A/ffiz), (4.3.8) for an extrinsic photoconductive detector, where another factor 2 has come in owing to the generation-recombination statistics of the carriers in the detector element (extrinsic semiconductor) [3]. Here, N should include both the photon rate N, from the source and the photon rate N35 from room temperature background in our calibration setup. From the equations (4.3.3) and (4.3.8), the noise-to-signal ratio (N/S) of the detector is calculated as: N 2m, + aw)“: ___ - . (4.3.9) S "IIZN' The equation (4.3.9) can be rewritten for the quantum efficiency: 4m. + N») n - . (4.3.10) Elias/8)2 If we measure N/S ratio of the detector, we can calculate the quantum efficiency from the equation (4.3.10). Noise and signal are measureduwith a PAR 5209 lock-in amplifier. The background photon rate N36 and the source photon rate N;.are calculated with the equation (4.3.6), using the following parameters: AD.“ - 3.32 x 10" (cmzsterad) for N. calculation; from the calibration setup An.“ - 2.04 x 10-2 (cmzsterad) for “so calculation; from the filter aperture of the detector with f/4.4 123 T3 - 572 K room temperature - 300 K filter frequency - 336 ~ 725 cm'1 filter transmittance - 0.72 window transmittance - 0.75 . The calculated and measured results are as follows: source photon rate ; N, - 3.13 x 101‘ (photons/sec) background photon rate ; Nae - 7.78 x 1015 (photons/sec) measured Gn ; Gn - 0.93 x 10'2 measured noise ; Nm - 3.9 x 10" (VA/il—z') measured signal ; Sm“ - 2.57 (V) The measured N/S ratio is 1.52 x 10*’(1//fi2) at the detector output, and the quantum efficiency n is thus 0.14 from the equation (4.3.10). It is obvious from the equation (4.3.9) that the N/S ratio will be much better if we reduce the background photon rate. The quantum efficiency n is in general a function of frequency and the value obtained above is therefore a kind of average over the relevant frequencies in the calibration experiment. The photoconductive gain G is not expected to depend on frequency since it has to do with the generated carriers in the detector element, but it does depend on bias voltage and its dependence can be measured. It needs to be corrected for other values of bias voltage. It is traditional to define the noise equivalent photon rate NEW as the photon rate that must be incident on.a detector for the output current to be equal to the RMS fluctuation per unit bandwidth in output current [3]. For a photoconductive detector this condition is Gne(NEN) - Ge(4nN)“Q. So the noise equivalent photon rate is: NEW - (413/:7)”2 . (4.3.11) Similarly, a noise equivalent power NEP can be defined as the incident signal power required to obtain an output signal equal to RMS output noise. The noise equivalent power is thus expressed as: 124 NEP - hcu(4N/n)l/2 . (4.3.12) Using these equations the noise equivalent photon rate and power are computed for our detector. The results are as follows: NEN - 4.8 x 108 (photons/sec//§E), and NEP - 5.1 x 10'12 (was), (4.3.13) where a median frequency v - 530 cmfl'is used for calculating NEP. Note that these numbers are measured in high background. They will improve in actual experimental conditions. The estimated NEW and NEP under actual experimental conditions are 3 x 107 photons/sec/AE and 2 x 10'13 W/fiE respectively (see section 4.5.2). The quantity Ga and the quantum efficiency 0 are needed to estimate the optical efficiency of the system and noise (see sections 4.4 - 4.6). As mentioned above the quantity Gn is in general a function of frequency and bias voltage. From the relations Vb“, - - Rf/RD Vu- (see equation 4.5.1.10 and Figure 3.6.1) and Vu- - RINemn) (equation 3.6.2), it is expected that the quantity 6» is linearly proportional to bias voltage Vb,“ as; Gr) - - [RD/(thfieflvbiu, provided that RD and R, do not depend on in“... Though the actual dependence is not linear (R0 is not constant) as measured by VI, versus Vbu“ the deviation is not significant for Vb“, - 2 - 6 V, so we have estimated Gn with a linear approximation for biases other than that measured. As for the frequency dependence, the measured 60 is a kind of averaged value over the spectral range in the calibration measurement, and is used for other frequencies with no correction. 4.4 Optical efficiency of the system The optical efficiency fun of the system is defined as the ratio of the number of photons per second arriving at the detector to the photon 125 rate from the radiation source accepted by the throughput of the system. The efficiency can be estimated from the expected losses of the optical components in the system. It can be expressed as: e... - R...” R. T. If T. e. . (4.4.1) where Ru is the reflectivity of a gold coated mirror, R, - 0.98; R, the reflectivity of the sample, R, - 0.60 (calculated value, see Figure 2.3.2); T the transmission coefficient of the polarizer, T - 0.7; T,, the P P transmission coefficient of a CsI window, T; - 0.8; T, the transmission coefficient of the filter, T, - 0.75; and ‘3 the grating efficiency, 5 - 0.8; parameters are obtained from the specification sheets Or a catalogs. Note that we consider only p-polarization. Similarly, we can define the optical efficiency of the spectrometer and transfer optics as the ratio of photon rate at the detector to that entering the spectrometer: e, - 12,5 1",, I, e, . (4.4.2) Substituting known parameter values into the equation.(4.4.l) and.(4.4.2), the estimated optical efficiency Gun of the system is 12.1 I, and the spectrometer efficiency 6, is 38.0 I for one polarization. 0n the other hand, we can determine the measured efficiency of the system from the output signal voltage and detector parameters. From the equation (4.3.4), the measured signal photon rate Nu“ at the detector is expressed as: 25%.“)... Nu... - . (4. 4. 3) GfleRIA‘ where (me)uu is root-mean-square voltage output, G the photoconductive gain, 0 the quantum efficiency, R, the feedback resistance andAt the gain 126 of the second stage amplifier. At 1000 cm'l, the measured signal voltage Wont)“, is 233 mV with 8.5 pm filter and the following parameters: Rf - 6 x 107 0, V1318! - 2.5 V, G17 - 4.2 x 10-3, As - 10, where Cu is determined by a linear extrapolation from the measured value at Vb“, - 6.0 V. The measured photon rate Nu,“ is therefore 1.6 x 1012 photons per second. Now the photon rate accepted by the throughput of the system can be calculated using the equation (4.3.6), with following parameters: Tum,“ - 1200 K, An.“ - 3.2 x 10'3 cmzsterad, v - 1000 cm'l, Au - 1.5 cm'l, e - 0.8 ; emissivity of the globar. The calculated photon rate No.1 is 4.5 x 1013 photons per second for p-polarization. Taking the ratio of the measured photon rate to the calculated one, we find the measured efficiency of the system: (steam. - 3.7 1. Compared with the expected efficiency of 12.1 1, ' approximately 70 1 of the radiation is lost due to alignment error, lower than expected component efficiencies, or other reasons we do not understand. Possibly, the quantity 0» is a lot smaller at 1000 cm'1 than at the lower frequencies (336 - 725 cm'l) where Gr) has been measured. 4.5 Calculated noise The required fractional noise level is S l x 10" 1M1; at frequencies less than 1000 cm'1 to study low frequency surface vibrational modes at submonolayer coverage. Possible sources of noise are the detector and its amplifier circuit, photon fluctuations from source and background, 127 instabilities of the source intensity and the chopper amplitude, sample movements, nonreproducibility of grating angles, movement of inner optical parts of the spectrometer relative to the shell, and noise from electrical measuring instruments. The expected noise of the system is estimated from two major sources; the noise from the detector circuit and photon noise from the system. In the following section, the calculated noise is compared.with the measured noise of the system. We have found that the dominant source of noise is scan-to-scan nonreproducibility. 4.5.1 Noise of the detector circuit A noise equivalent circuit of the detector and preamplifier is shown in Figure 4.5.1 [4]. The various noise sources represented in this model are: eG - voltage noise of JFET input stage, is - current noise of JFET input stage, if - Johnson current noise of the feedback load resistor, in - Johnson current noise of the detector. R: is the feedback load resistance with its capacitance C: and R0 the detector resistance with capacitance CD. Note that the JFET, the feedback load resistor and the detector are at liquid helium temperature, as shown in Figure 3.6.1. Let the complex impedances of the feedback load resistor and the detector be 2, and 20 respectively, then 2, and Zn are: R: z, - , (4.5.1.1) 1 + inoctllf RD zD - . (4.5.1.2) 1+imCDRD 128 CD consists of the dielectric capacitance Cd“ of the detector and JFET gate capacitance (33' Cd“ is calculated with the formula: Cd... - eA/d, where e is the dielectric constant of the detector, A the cross-sectional area, and d the length of the detector. Using 6 - 11.7, A - 4.0 x 10'3 cm2 and d - 0.1 cm, Cm“ is 0.52 pF. The JFET gate capacitance C8 is 4 pF, from the specification sheet. CD is therefore 4.52 pF. C: can be estimated from measurements of the voltage gain of the circuit with an AC signal injected at the bias point, see Figure 3.6.1, as a function of frequency, using gain - - Zf/ZD. C, is approximately 0.5 pF. The roll-off frequencies (21rRC)"1 for R, and the detector itself are 5.3 K32 and 2.5 KHz, respectively. At a chopping frequency of 800 Hz, they are basically resistive, though the detector combined with the gate capacitance of JFET will behave capacitively. Under operating conditions the detector is under constant voltage bias, so its capacitance does not affect circuit response. We assume that the noise sources are independent, so that we can analyze them separately and find out their contributions to noise at the input test of the circuit. The amplifier gain a is assumed to be sufficiently large throughout the noise analysis that the amplifier input is a true virtual ground. The noise at the input test point due to the voltage noise of the JFET input stage is: I 20 + Z, Wyn)“- - —— es . (4.5.1.3) Zn The noise current of the JFET input stage flows through the load impedance to appear at the input test so that the noise is: 129 (vgcn)II ' 1922- (4.5.1.4) Combining these two noise sources, the noise due to the JFET input stage is expressed as: (vfin)Ir - ((1 + z,._./z,,)2e,,2 + (15292)“? (4.5.1.5) The Johnson current noises of the detector and the feedback load resistance are expressed as: 1,, - (4kT/RD)1/2 , (4.5.1.6) and if - (4kT/R£)1/2, (4.5.1.7) respectively. Similarly to the current noise of JFET input stage, these current noises appear at the input test as 102, for the detector resistance and 1,2, for the feedback load resistance, respectively. The noise at the input test due to these two sources is expressed as: (an)Ir - (4kT(1/RD + 1/R,))1/zz, . (4.5.1.8) The total noise (Vnnhr due to the detector circuit is a quadratic addition of these noises: Wan)“ - ((V‘“)r,2 + (VJ“)I,2)1/z. (4.5.1.9) Substituting some typical numbers, we can estimate the predicted noises from the various noise sources. The ratio of the detector resistance to the feedback load resistance under operating conditions is measured from the dc input test and dc bias voltage of the detector circuit: Rt/Rn " ' Vines/VI! - (4-5-1-10) Using the bias voltage and measured input test voltage when the spectrometer, transfer optics and the detector are cold, Rt/RD - 0.30. Since Rf - 6.0 x 107 0, RD is 2.0 x 108 0. The upper limit of the voltage 130 and current noises of the JFET input stage are 2.0 x 10'8‘V//fiz and 6.4 x 104‘ A/JEE respectively, from the specification sheet. From equations (4.5.1.1) and (4.5.1.2), zq/ZD z 0.31 + 0.091 at 800 Hz. The estimated noise from the JFET input stage, which is at 4.2 K, at the input test is thus (using the equation 4.5.1.5): (v83)IT - 2.6 x 10'8 V//H—z. (4.5.1.11) From the equation (4.5.1.8), the Johnson noise from the detector and feedback load resistance, which are both at 4.2 K, at the input test is: (thr - 1.34 x 10'7 V//H_z', (4.5.1.12) where the Johnson noise due to the feedback resistance is actually dominant. Adding these two noises quadratically, the total noise due to the detector circuit is: (an)IT - 1.37 x 10'7 V/lfi. (4.5.1.13) From this estimation, we expect the Johnson noise to dominate the detector circuit noise even though the feedback load resistor is cooled to liquid helium temperature. In fact, room temperature feedback resistors were used in the original detector amplifier circuit, which has been modified to include a cold preamplifier and cold feedback resistors to reduce the Johnson noise. Detector circuit noise is now'much smaller than the background photon noise (see below). 4.5.2 Photon noise The expected photon noise can be calculated for background limited operation of the detector. The noise current in the detector due to the fluctuations of incident photons is given by the equation (4.3.8). The noise current flows through the feedback resistor and appears as a noise voltage at the input test. The noise voltage is therefore expressed as: 131 (Vp“)n- - 2Ge(nN)1/ZZ£ . (4.5.2.1) Substituting some typical numbers, we estimate the noise voltage at the input test to be: (Vp“)n- - 1.4 x 10" V/JEE ; from 100 K background, and (Vp“)n- - 2.7 x 10'7 V/fi ; from the source photon rate. The background photon rate N36 and source photon rate NS used to estimate the photon noise are: N36 - 4.2 x 1013 photons/sec NS - 1.6 x 1012 photons/sec , (4.5.2.2) which are calculated using the typical parameters: C - 3.0 x 10'2 n - 0.138 thRt-6x1070, A0,“ - 8.5 x 10‘3 cmzsterad r - 1.0 for the background calculation, T36 - 100 K V - 350 - an AD.“ - 3.2 x 10'3 cmzsterad r - 0.03 for the source photon rate Tum,“ - 1300 K calculation. Au - 2 cm'1 11 - 1000 cm'1 Background photon noise is much bigger than source photon noise, and is still dominant, even though the spectrometer and relevant optics are cooled to 100 K“ This is mainly due to the fact that the background noise is integrated over frequency with efficiency 1 - 1.0, while the signal is monochromatic, with efficiency'r - 0.03 that comes from the product of the total efficiency of the system fun - 0.037 and the globar emissivity e - 0.8 (see section 4.4). Note that the blackbody radiation of 100 K has a peak at about 250 cm and that the detector has a low frequency cut-off, 350 cm“, (see Figure 2.6.1). All the calculated noises are summarized in 132 Table 4.5.1, along with the measured noises. 4.6 Measured noise 4.6.1 Noise from the detector circuit As discussed in section 4.5, the noise at input test consists of noise from the detector circuit and that from photon fluctuations if other noise sources are not significant. If we block the entrance aperture of the detector by a metal plate that is cooled to liquid helium temperature and make photon noise negligible, the noise due to the detector circuit can be estimated from the measured noise at the input test. The noise is measured either by an HP spectrum analyzer, or by a computer with a spectrum analyzer program, a preamplifier and a band pass filter, or by the lock-in amplifier. They give consistent results. The measured noise at the input test is: (Vr“)u - 2.8 x 10’7 WIFE, (4.6.1.1) with Vb“. - 3.0 V and VI, - 44 mV. Since the photon noise of 4.2 K background is negligible, the measured noise should mainly come from the detector circuit. This noise is larger than the estimated one, 1.4 x 10'7 V/Jfiz, by a factor of two, possibly because the detector is under voltage bias, 3.0 V, giving rise to excess noise. 4.6.2 Photon noise It is not easy to isolate photon noise from noise of the detector circuit in measurements, but we can measure the photon noise just by measuring the total noise at input test voltage since background photon noise is expected to be dominant. The photon noise can.be estimated also from the measured input test voltage, since that voltage is proportional 133 to the total photon rate (see equations 3.6.1 and 3.6.2). Using equation (4.5.2.1) and the relation Vn- - ZteGnNBG, the photon noise is expressed as: (Vp“)n- - 2(ceztvu)1/2. (4.6.2.1) When the spectrometer and the transfer optics are cooled by liquid nitrogen, the measured input test voltage Vu- is -0.76 V at the bias voltage Vb,“ - 2.5 V and frequency v - 1000 cm”. The estimated photon noise from the equation 4.6.2.1 and measured Vn- is therefore 9.34 x 10'7 V//H_z. The calculated photon noise is 1.39 x 10" V/flE, which is slightly higher than the estimated noise from the measured input test voltage, indicating that the‘ actual background level is slightly lower than expected, but lower by about 20 1 than the noise measured with a spectrum analyzer, 1.8 x 10'6 V/flE. We have measured the photon noise from the room temperature background when the spectrometer and transfer optics are warm. The measured photon noise is 1.08 x 10‘5 V/JIE. The estimated noise from the input test voltage is 9.9 x 10'6 V/fil-Z and the calculated photon noise from room temperature background is 1.0 x 10'5 V//ll_z. The agreement is very good. Maybe, the agreement for the room temperature results is better, because the G!) values are measured for the characteristic frequencies of 300 - 500 K blackbodies. For the purpose of comparison, we have also measured the photon noise from the gate valve between the transfer optics and detector. The input test voltage is -9.75 V at a bias of 2.42 V. So the estimated photon noise from the input test voltage is 3.4 x 10" V/JFIE. The measured noise at the input test is 2.8 x 10'6 V/fi, which is in good agreement with the estimate. But this is smaller than that of room temperature background by 134 a factor of three, suggesting that the 4.2 K ambient radiation of the detector cryostat is reflected back to the detector from the gate valve. With the cold spectrometer and the room-temperature transfer optics, the measured noise is 2.1 x 104’V//fiz, which is larger by 50 1 than the calculated 100 K background noise. The extra noise most likely comes from stray radiation. This is the reason why we have modified the transfer optics to be cooled by liquid nitrogen, and installed liquid-nitrogen cooled baffles in the optical path both between the spectrometer and the transfer optics, and between the transfer optics and the detector. With this modification, however, the measured.noise decreases only by 10 2, not so much as expected, by a factor 2, from the measured input test voltage. This discrepancy may come again since we use 00 values measured for 300 - 500 K blackbodies, as mentioned above. 4.6.3 Behavior of system noise As we have discussed in section 4.5, we expect the noise at the detector output to be background photon noise limited. From the equations (4.5.2.1) and (4.6.2.1), the noise intensity (Vp“)n- should have'a square- root dependence on the input test voltage Vnror the photon rate N at the detector. A log-log plot of (Vp°)n- versus V“- or N should show a straight line with a slope of 1/2, if the detector is operated under the photon noise limited condition. We have measured the noise at the input test as a function of input test voltage, which is shown in Figure 4.6.1, with a bias voltage of 10 V. The photon flux, and therefore the input test voltage, are varied by rotating the liquid-helium-cooled filter wheel in the detector. The noise shows a photon noise limited behavior, except near V11- - 10 V. 135 1 we measure a signal level of ~ 270 mV At a frequency of- 900 cm' with a source temperature of 1300 K. Using the signal value and the noise ' numbers we have discussed so far, we can estimate the sensitivity for each case. First, the noise-to-signal (N/S) ratio would be 1.0 x 10‘5 l//H_z if we were limited by source photon noise. This is the best we could possibly do with the present source, detector, and the optical throughput and efficiency of the system. We have not achieved this level of sensitivity. Second, the background-photon-noise limited N/S ratio is 6 x 10'5 l/flE. This is the best our current instrument could do with perfect reproducibility, and is what we currently seek to achieve. This level of N/S ratio is routinely achieved in time scans at a fixed grating angle. A noise spectrum measured at the input test is shown in Figure 4.6.2. The noise spectrum has a l/f noise feature at low frequencies, and a long tail extending to 800 Hz. We have chosen 800 Hz as the modulation frequency to avoid this noise. Tuning fork choppers have a tradeoff between high frequency and high amplitude. It is related to the maximum stress that can be applied in the vibrating arms. If we tried to go to higher frequency we would not have enough amplitude. There are no sharp features due to vibrations in the noise spectrum, which means that the system is very stable and insensitive to the mechanical vibrations. So far we have limited our discussion to the noise with no grating movements involved. In real infrared spectroscopic measurements, we need to compare a spectrum taken with adsorbates on the surface with one of a clean surface. In order to measure the noise that matters in the real surface spectroscopic measurements, we need to take and compare two 136 sequential spectra with clean sample surface. In this way we have reached a noise-to-signal (N/S) ratio of 2 x 10" 1/fil—z' near 850 cm'1 and the extra noise comes mostly from non-reproducibility of one scan to the next. In Figure 4.6.3, two noises are shown; one from a time scan and the other from a real frequency scan. In the real scan noise, the spikes from the non-reproducibility dominate the noise. This is discussed in the next section. 4.7 Improving noise-to-signal ratio In addition to the noise sources we discussed in section (4.5) and (4.6), namely detector noise and background noise, there are other possible noise sources in the system. In this section I describe those noise sources and other efforts that we have made to improve the N/S ratio of the system. 1) Instability of the radiation source can be caused by instability of the power supply, or mechanical movements of internal optical parts relative to the external optics. 2) Instability of the chopper or chopper controller can be a problem. While phase jitter is the only problem of the conventional rotating chopper, amplitude stability is crucial for a vibrating chopper. The fractional noise of the chopper can be measured by a lock-in amplifier feeding the controller signal as an input and reference signal. We have modified the circuit of the chopper controller to improve the fractional noise from 2.0 x 10" l/JHE to 2.0 x 10‘5 1//sz:. At this level it is well below background photon noise. 3) Though mechanical vibration of the sample does not contribute directly to the noise because of chopping, it can affect the 137 reproducibility. The sample is required to travel more than 8 inches in the sample chamber between the IRAS level and the surface analysis level. The total length of the sample mount is 30 inch long which is made of 20 mil thick stainless steel tube. The manipulator is a cantilever model which has three lead screws for better stability. We know that l) - 3) do not dominate because they would show up in time scans. The dominant noise comes only when we scan the spectrometer. 4) Non-reproducibility of the optical efficiency of the spectrometer can be a noise source. The optical efficiency of the spectrometer can change from scan to scan due to mechanical motion of the internal spectrometer optics relative to the external cryostat, which may be caused by the grating drive force or by the vibration of vacuum pumps. As discussed in section 3.4, three clamping screws have been introduced through the walls of spectrometer Chamber'to keep the spectrometer optical table from moving around. Also the grating can tilt. A tilt of a few x 10'5 radians can shift the image at the exit slit vertically by ~5 x.10“ of its height. We have modified the grating mount and drive to include a pair of precision bearings at the worm shaft of the grating table and a pair of angular contact bearings at the worm driving shaft, to achieve more mechanical rigidity. 5) The dominant noise source turns out to be the non-reproducibility of grating angle and/or efficiency. This noise source appears only when two spectra are measured and compared each other. In each spectral scan the signal intensity is measured as a function of grating angle over a range of angles. If the angles in one scan are not exactly reproduced in the next scan, the result appears as noise in the comparison of two 138 spectra. Suppose that instead of returning to a particular angle 0 on the second scan, the grating returns to a slightly different angle 0 + A0. Since the light intensity I reaching the detector depends on the angle, the intensity measured in the second scan is different from that in the first by AI - (dI/d0)A0. Since A0 varies randomly from point to point in the spectrum, these angular errors appear as signal fluctuations when the two spectra are compared (see section 3.4.3). It is expected that the noise is greater as the slope of the intensity is larger, which is exactly observed. We can estimate the non-reproducibility of the grating angle from the noise. As discussed in 3.4.3, the original grating drive was a worm-to-worm gear combination with a 360-to-l gear ratio and a bellows coupling to a stepper motor. The angular non-reproducibility was approximately 2 arc sec with the N/S ratio of 4 x 10" l//Hz around 850 cm“. This noise cannot be eliminated by any modulation, and was not improved.by averaging scans since it was not random. We made many efforts to enhance the mechanical rigidity. These efforts did help, but were not enough. We decided to install a high-resolution angular transducer and a feedback control system (see section 3.4.3) and the N/S ratio has been improved to 2 x 10" l/flE around 850 cm'l. But the nonreproducibility of the grating angle and/or efficiency from a scan to the next'is still the dominant noise source. We have installed a half height slit in the entrance slit wheel to investigate the problem. Since the system noise is limited by the scan-to-scan nonreproducibility, the system noise is linearly proportional to the 139 signal level. That implies the following: a) increasing the signal level does not help - brightness of the source is not a limitation, b) further reducing background photon noise does not help - it is not a limitation, either, c) FTIR would not necessarily have the multiplex advantage [3]. We are exploring several possible avenues to improved sensitivity: a) further improvements in the rigidity of the grating mount and drive, b) double modulation, combining intensity modulation with wavelength modulation. Higher signal intensity may be needed to take advantage of this technique, c) use of a detector array to obtain true multiplexing and avoid the necessity of scanning the grating. 4.8 Comparison of our system with others When we started our project in 1988, no IRAS systems were available for the spectral range below 800 cmPK. Since then, however, several groups have developed IRAS systems which extend the spectral range into the far- infrared region (as low as a few hundred cm'l), maintaining a reasonable sensitivity. Table 4.8.1 summarizes some features of those systems. While two groups, the Brookhaven National Laboratory (BNL) group and the Bermudez group, are using brighter sources to improve the N/S ratio, others are using conventional blackbody sources, such as globar sources, and room temperature spectrometers, mostly FTIR. Using a vacuum FTIR and synchrotron radiation as a radiation source, the BNL group has, on rare occasions, been able to achieve a N/S ratio of 6 x 10'5 in a frequency region around 100 - 800 cm'1 [5]. Synchrotron 140 radiation instabilities dominate the system noise and a more typical N/S ratio attainable is ~2 x 10". They have measured the metal-carbon stretching band for CO on Cu(111) and Cu(100). Bermudez et a1. [6] have developed a diode laser IRAS system using a ZnSe photoelastic modulator (PEM). Two diode lasers are used to cover the spectral range from 600 cm'1 to 1020 cm'l. Spectra are measured without a monochromator by varying the diode temperature. The resolution is limited by the energy range over which multimode laser emission occurs at a given temperature and current, and is ~10 cmfi'at best. The best N/S ratio is ~3 x 10“,‘which is seemingly limited by the lasing instability. Later, they were able to obtain a N/S ratio of ~1 x 10" in a frequency range 600 - 1000 cm'l, using an FTIR spectrometer with a HngTe detector and likely a conventional globar source [7]. Ryberg [8] has measured the metal-carbon stretching band for CO on Pt(111), using a vacuum grating spectrometer with wavelength modulation and a W - filament source. Apparently he has achieved a N/S ratio of l x IIY‘ or better, and likely background photon fluctuation noise dominates the system noise. According to our calculation, the source temperature should be higher than 1500 K in order to achieve that N/S ratio. The Trenary [9] and Bradshaw [10] groups have developed very similar systems and measured the metal-carbon stretching band for CO on Pt(111). They use vacuum FTIR's, globar sources and similar sample sizes. The N/S ratios of these systems are likely limited by background photon fluctuation noise. The Greenler group [11] has developed an IRAS system using a grating spectrometer, globar source and thermocouple detectorz They'have measured 141 the Ag-C stretching mode of C0, and the 0-0 and Ag-O stretching bands of oxygen on a polycrystalline silver foil. Considering that they are using a ribbon sample which allows them to use a large sample and therefore large throughput, the fractional noise of the system is not so impressive, ~3 x 10" (estimated from the published data), and is probably limited by detector noise. To summarize, two groups, BNL and Bermudez, use brighter infrared radiation sources to improve N/S ratio to overcome the background photon fluctuation noise, while four other groups use conventional blackbody radiation sources and must have developed very stable systems, mechanically and. optically, so that the noise-to-signal ratios are approaching the background-photon-fluctuation noise limit. However, the former two groups have noise sources from the source itself, limiting the N/S ratio, and the other groups are expected to have intrinsic difficulties to further improve N/S ratio from the background-photon- fluctuation noise. 0n the other hand, we have designed and built our system to minimize the background photon fluctuation noise by cooling the spectrometer and relevant optics. The present N/S ratio of our system is 2 x 10" l/JFE and is limited by nonreproducibility of grating movement and/or spectrometer efficiency. If we improve the reproducibility of the spectrometer, the N/S ratio can be improved by a factor of 4. 4.9 Summary We have measured and analyzed the performance of the system. Most of the characteristics, such as the spectral accuracy, the frequency resolution and optical efficiency, are satisfactory, but the sensitivity 142 does not completely meet the design target, 51 x 10" 1M? The sensitivity of the system achieved to date is 2 x 10"ZL//HE at ~850 cm‘l, which is good enough to measure the external and internal vibrations of many adsorbates on metal surfaces, but needs to be improved for the study of other weak external modes such as atomic adsorbates. The sensitivity is limited by scan-to-scan nonreproducibility, rather than random noise. If we measure noise at a fixed frequency, therefore at a fixed grating angle, as a function of time, the fractional noise drops to 6 x. 10'5 l/flE. This noise comes mostly from 100 K background radiation. The measured detector noise at input test is 2.8 x 10'7 V/flE, which is not significant compared with the 100 K background noise, 1.8 x 10" V//HE, as expected, but is higher than the calculated noise by a factor of 2 probably due to the bias voltage. The source photon noise estimated from the signal intensity is 2.7 x 10'7 V/flfi, which is not significant either. The sensitivity of our system can be improved to a level of 6 x 10'5 l/JHE if we improve the reproducibility of the spectrometer. There is no far-IRAS system that routinely operates at this level of sensitivity. The IRAS apparatus at Brookhaven National Laboratory has a sensitivity of approximately 2 x 10" l/JITE, which is apparently limited by the stability of the synchrotron radiation source, while others have the room- temperature-background limitations. Other possibilities for overcoming non-reproducibility include wavelength modulation or using a linear array detector. 143 References 1. 10. ll. P.R Bevington, Data Reduction and Error Analysis for the Physical Sciences, (McGraw-Hill Book Company, New York, 1969). S. M. Sze, Physics of Semiconductor Device, 2nd edition (John Wiley & Sons, Inc. New York, 1981) pp. 744 - 748. P. L. Richards and R. G. Tobin, in Vibrational Spectroscopy of Molecules on Surfaces, edited by J. T. Yates, Jr. and T. E. Madey (Plenum Publishing Co. New York,l987) pp. 417 - 463. E. L. Dereniak, R. R. Joyce and R. W. Capps, Rev. Sci. Instrum. 48, 392 (1977). C. J. Hirschmugl, G. P. Williams, F. M. Hoffmann and Y. J. Chabal, Phys. Rev. Lett. 65, 480 (1990). V. M. Bermudez, R. L. Rubinovitz and J. E. Butler, J. Vac. Sci. Technol. A6, 717 (1989). V. M. Bermudez and A. S. Class, Langmuir 5, 316 (1989). R. Ryberg, Phys. Rev. 40, 5849 (1989), Phys. Rev. 40, 8576 (1989), Phys. Rev. 40, 10273 (1989), Surf. Sci. 114, 627 (1982). I. J. Malik and.K. Trenary, Surf. Sci. 214, L237 (1989); V. K. Agrawal and M. Trenary, J. Chem. Phys. 95, 6962 (1991). D. Hoge, M. Tflshaus, E. Schweizer and A. M. Bradshaw, Chem. Phys. Lett. 151, 230 (1988) X.-D. Wang, R. G. Greenler, Surf. Sci. 226, L51 (1990), Phys. Rev. 43, 6808 (1991), X.-D. Wang, W. T. Tysoe, R. G. Greenler and K. Truszkowska, Surf. Sci. 257, 335 (1991), Surf. Sci. 258, 335 (1991). 144 Table 4.3.1 Detector Calibration The responsivity of the detector has been measured for three different temperatures, using the calibration setup shown in Figure 4.3.1. efficiency of the detector optics is included in the responsivity and the amplifier gain A, used is 10. The Ts (Avout)pp Rf Pchop N chop 3 3’7 (K) (V) (K0) (#W) X101‘(S'1) (A/V) X104 380 0.67 500 0.88 0.83 0.15 1.01 518 0.37 100 2.67 2.48 0.14 0.93 573 0.47 100 3.45 3.20 0.14 0.92 145 Table 4.5.1 Summary of calculated and measured noises. See sections 4.5 and 4.6 for details. Calculated Estimated Measured noise noise (V//HE) noise from Vfr at input (V/ffi) (VA/HE) Detector circuit 1.4 x 10'7 N.A. 2.8 x 10"7 noise Photon noise: 1.4 x 10" 9.3 x 10'7 1.8 x 10'6 100 K background Photon noise: 300 1.0 x 10'5 9.9 x 10‘6 1.1 x 10'5 K background Photon noise: 1300 2.7 x 10'7 ...................... K source 100 K plus 300 K .......... 2 3 x 10'6 2 l x 10"6 BC from transfer optics 146 .003 huwmcouca Hvac comm - omm uouospcooouona 00H000-~zq \.uooam \7588 L: x N .5 a x 3 m .3 93 -- 55on 838 636385 pm: 7.8 83 - 8... 56:6 Ea: 5E 6.66% E \7500me .79." x m 0.20 30 GA .0» Home." 00030 nouns—08:00:06 0: woman—yon Tao ooom - cos . \Taooome .Toa x H 330 uouodpaooouoza .8.»qu 9830.: .3 Taoooouw Pea x m .020 .8 m6 2% .0080 unnoaw \MHE 850a> 30£mvoum 7.5 83 - 2: 8.5:. m< .62.. .53 63:63 Hmummuohaom omuam was >80 _oH_ \9roH x ml Sow x a0~.H w oaasoooauocu Hugoam ..uoomm wcwueuw Hoacoouu .780 aces - coo poaooo xmwc .mwp H0003 Houuwa wca>0a _m. \RioH x Hz 80 m~.H .ww Houoaoaon um -- Honoaw \MHBh assoe> huocouh dTao comm - own Aevxmfip uouodpcooouosn. unoaoawm .006 SuwSoH0>m3 _n_ \Rioa x at .mwp 30 H .m.um sonou - 3 \wcHumum azaou> wuonxm dv.8 ooos - oma nouuwe wcu>0a \Smoacfivfg x N 363 8.08 \Eou 038:3 7: 9.8» 3.6.05 Tea x w .8 H x 30 N .mw uouoaoaon muwm 80.89.3083 MHE 8508' A.z.m owcou hocozvoum oocopwocw oouaom soauaascoa \0Humu m\z ouum mamamm m0 mamas 80000000 uswaa \uouolouuooAM .mmwoc vasoumxomn weapon 00 mouuao 00m Houoaouuoomm 00H000-:ow0uuwc-0wscaa 0 magma mg sz .cowuauomne uwuonmnoaum onu weapon ou assom> 0H 00008090 haumoa .uouoaouuoomm eunumuoaaou-aoou 0:0 moousom HoaHoSu Hoaoauce>800 mafia: one muonuo magma .xuwmsoucw Hmawam 0:0 0>0HQEH 00 moouaom Houswwun mafia: one .9008& nopdsuon 0:0 naouw qzm .masouw 039 maoumam m5: m0 803.8980 H.m.¢ OHQMB 147 tic axis / °p light in light out \ grating normal 6 \ We grating angle Z crating Figure 4.1.1 Light at oblique incidence onto a grating, showing the . definition of the angles ¢ and 0. 148 (Hm) n3. ' 0.0 1.0 2.0 3.0 4.0 5.0 2dsinO (le) Figure 4.1.2 Calibration of spectral frequency for cold spectrometer. The calibration curve for room temperature spectrometer is not shown since it is indistinguishable. A nominal value of d is used for 2dsin0. 149 1]? 0.006 m r l t l l I 'E 02mmfln 5 ......... 0.5mm . 0.005 F / ....... - 1.0mm _ _o , ------- 1.5mm S... / ——- 2.0mm 3 / >. 0.004 — / i - f. I i {3 I :' 3 0.003 — I :' — .5 I : I :' 0.002 » / - . / f / I 0.001 1 / f . f 3. ~ / I” I. ‘ . {'l’ ,I'A f ~\‘ 0.000 “2,481.34 A~ '1 . JAM— -0.15 -0.10 -0.05 0.00 0.05 0.10 0.15 Angle (degrees) A Figure 4.2.1 Measurement of the angular resolution of the spectrometer. Spectra of a He-Ne laser are measured for different slits. Peaks are not centered at zero degrees The since the grating angle was not calibrated when the measurements were made. intensities are arbitrarily scaled for convenience. 150 0.10 - — Calculated - 7; 0.08 - — Q) Q) ,- e 80 -".. Measured m 0.06 — -+ 3 2 . E 0.04 ~ ,s’ ' - a. 0.02 — - -_ r . ' 0.00 l l A l i l l l l 0.0 0.5 1.0 1.5 2.0 2.5 Slit Width (mm) Figure 4.2.2 Measurement of angular resolution of the spectrometer. Both calculated and measured values are shown. The measured resolution behaves as if the effective focal length were 74 cm instead of 65 cm. 151 /— LHe Source aperture A 2 ]/ /’ Window / Blackbody source i” a / / L6 d / Spectral filter / with aperture A I — Winston cone Chopper wheel / L Detector Figure 4.3.1 Calibration setup for the detector. The distance between the source aperture and the detector is 10.9 cm. The temperature of the blackbody is measured with a K-type thermocouple attached to the copper cone. 152 Input Test ,\ . / O Bias Test Figure 4.5.1 Equivalent circuit of detector and preamplifier 153 (V6 ”)0 (V/VHZ) l0‘6_ l.l.l llllllll.LA¥J 1 1111111 10“ 100 101 V|T(V) Figure 4.6.1 Measured noise as a function of input test voltage. The line has a slope of 1/2 indicating the photon noise limited operation of the detector with 11 = 0.1. 154 .8085? =88:er 3 0.6 3.238» 82” 0: can 205 “m5 302 .30 $55 mm 80: e: 20:3 .Nm 8w 3 88:0 mm accuses—a. ”530:0 05. .t0 880:0 05 53, :5 sweets: 2%: 858 55 “m3 “:9: 05 an 8388 858% came: 4.. ~04 8:me Anzv accesvosm 000—. can.— 000. 000 3.0 _ _ _ _ a _ 4. 00.0 I ........... n ............ Mo ...................................... u ............ I ....................... l j ........... . ............ . ........................ t ........................ l ON.O I. ............................................................... W ...................... I1 0'0. 0 m / m s I ........... c ............ o ...................................... m ............ I oooooooooo l o m u m i I ........... . ............ . ....................................... . ............ t ....................... l 0%.. ) m H m z .I ........... . ............ s ...................................... m ............ s ..................... nl ( I ............................................................... m ................................ n] 00.. r ........... . ............ . ..................................... ............ . ................... l m l x r _ _ _ r _ r n u 155 1.5E-3 T T fl f . T ' f . I . I T T l — time scan 0 1013-3 ‘ . --------- real scan ‘ <1 : r. 5.0E—4 ~ g , ‘ 3: ~53 ; 555: , 5'; 3. .l ‘ :: : ' ‘ :0. :3 '; :31. :2 o ::.:'.: i 0... .0... q. ’0... 0.0EO - 3.1 .. 5;; . " .::..-°' : - _ —5.0E-4 r i i _10E_3 . l 1 ~. 1 . l . l l l . 780 800 840 860 880 900 920 940 frequency (cm-1) 820 Figure 4.6.3 Noise from a ”time scan” and ”real scan". The frequency label of the x- axis applies to the real scan. It takes approximately 3 minutes to measure the time scan. Deviation from the mean value is shown for the time scan and a comparison of two subsequent spectral scans is shown for the real scan. 5. Measurements on Oz/Pt( 111) and CO/Pt(111) In this chapter, I present vibrational spectroscopic measurements of - CO on Pt(111) and 02 on Pt(111). They have been widely studied, partly due to the practical application to catalysts for oxidation of carbon monoxide. In particular, carbon monoxide on Pt(111) has been extensively studied by both infrared and electron energy loss spectroscopies. But infrared measurements of low frequency modes, O-O, C-Pt, and O-Pt, are very limited despite their scientific as well as technological importance. We have chosen CO on Pt(lll) and 02 on Pt(111) as the first two systems to measure by our new IRAS system. 5.1 Sample preparation and characterization The single crystal platinum sample was purchased from Aremco, Inc. The original sample was approximately rectangular, with dimensions 3 cm wide x 1 cm high x 0.1 cm thick. We cut off the sample edge at one corner by 0.3 cm x 1.5 cm in a triangular shape because of misorientation by approximately 9° from the [111] direction while the other region is misoriented by 2.7°. The sample was repolished mechanically with 0.5 micron diamond paste. The normal to the sample face happens to be oriented at 2.7° from the [111] direction, as determined with the back reflection Laue X-ray technique. The crystal was examined with low energy electron diffraction (LEED) . No crystallographic faults could be detected with LEED. Figure 5.1.1 shows a LEED pattern with a doublet that is from the ordered step structure. Since the intra-doublet separation is inversely proportional to the distance between steps, the terrace width can be obtained by 156 157 comparing the doublet separation with a major spot separation. The terrace is 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 [2]. For the {00} spot, the voltages Vm,(in.Volts) at which a singlet spot exists with maximum intensity are given by the equation [2,3]: 150 V°o(Singlet Max.) - S2 , (5.1) Adz where d is the step height in A and S is an integer. For half-integer values of S doublet spots of equal intensity exist. Table 5.1 shows our measured values of V°o(Singlet Maximum) and V°°(Equal Doublet), and estimated step heights in A. It is clear that the steps are monatomic, since the expected step height is 2.12 A (the radius of a Pt atom is 1.221A). From the terrace width and step height, the normal to the sample surface is at 2.73° from the [111] direction, which is in very good agreement with that determined from Laue X-ray diffraction. The primary contaminants in the crystal were carbon and calcium. The calcium was removed by argon ion sputtering. Sputtering is performed at room temperature for about 10 - 20 minutes with Ar pressure at 5 x 10'5 Torr. The sputtering energy is 400 eV and the ion current reaching the sample is approximately 5 nA. The sample is then annealed at 1000 K for 30 minutes after sputtering. The carbon is removed by heating the sample in oxygen (1100 K, S x 10’8 Torr with the closer) for several hours, and the sample is then heated to 1375 K for 10 - 30 minutes to remove residual oxygen on the surface. As the carbon level decreases to less than 102 of a monolayer, lower oxygen pressure (1 x 10’8 Torr) is used for removing 158 carbon. The last few percent of a monolayer of carbon are removed by dosing the sample with 0.5 - 1.0 L of oxygen through the closer [I Langmuir (L) - 1 x 10'6 Torr for 1 sec. , the closer enhances the effective dosing by a factor of ~20 ; see section 3.3.4], and then heating the sample to 1375 K for S - 10 minutes. During the oxygen treatments, calcium often comes out of the bulk and is removed by argon ion sputtering. Sputtering always results in new carbon contamination, and additional oxygen treatments are necessary in turn. After numerous cycles of cleaning, residual concentrations of both carbon and oxygen could be reduced to $3 1 of a monolayer (noise level). Figure 5.1.2 shows an Auger spectrum of such a surface. About 2 - 3 percent of carbon usually appears in the Auger spectrum and is thought to be picked up from the Auger filament. 5.2 Measurements of vibrational spectra for CO on Pt(11 1) In this section, we present measurements of the C-0 and Pt-C stretching modes for CO on Pt(111). Carbon monoxide on Pt(111) is one of the most extensively studied chemisorption systems by a wide variety of experimental surface techniques. There have been reported numerous studies using infrared reflection absorption (IRAS) [3-8] and electron energy loss spectroscopy [9-13] ,as well as low energy electron diffraction (LEED) [4,12], temperature programmed desorption (TPD) [4,5,12], ultraviolet photoelectron spectroscopy (UPS) [l3], and inelastic helium scattering [I4] . With such a broad experimental background carbon monoxide on Pt(111) could serve as a reference system. It might be natural that CO on Pt(111) is chosen to see the performance of our system. On the Pt(lll) surface CO is adsorbed non-dissociatively with the carbon bonded to the surface in a linear (on-top) or a bridged 159 configuration depending on the coverage. At a fractional coverage 0 S 0.33 and temperature of 150 - 170 K, a diffuse (fix/3)R30° overlayer structure is reported, which with increasing coverage is continuously transformed into c(4x2) at 0 - 0.5, and then into (fIV—ZxfififlllT overlayer structure at 0 - 0.58 [12]. In our experiment a sharp c(4x2) pattern is observed at 90 K with 0.5 L of C0 exposure with doser. The sample is dosed with C0 gas using the closer at 90 K and annealed at 290 K for 3 minutes and then cooled back to 90 K. This observation of a sharp c(4x2) pattern means that the sample cleanness is well controlled. In temperature programmed desorption (TPD), one main desorption maximum appears between 400 and 600 K. Figure 5.2.1 shows a TPD curve for a CO exposure of 0.5 L. Note that CO desorption around 650 K is negligible compared to the main peak. It is generally believed that desorption above 600 K is due to desorption from imperfections like kinks and steps. Therefore the defect concentration on our (111) surface is considered rather small, but we know it is at least 5 1 from the LEED analysis. IRAS spectra are shown in Figure 5.2.2 for the C-0 stretching mode of on-top CO on Pt(111) with different exposures and in Figure 5.2.3 for the Pt-C stretching mode. The spectra are measured using the room-temperature spectrometer with cos¢ - 0.987 for the C-0 mode and using liquid-nitrogen- cooled spectrometer with cos¢ - 0.982 for the CoPt mode, respectively (see section 4.1). The absorption intensity and linewidth of the on-top C-O band are about 6.5 2 at 2109 cm'1 and 3.8 cm'1 at the saturation coverage, measured with instrument resolution 2.9 cm'l. Compared with other works, these 160 values are in good agreement; Luo [15] has reported an intensity of 5.5 ~ 7 2 at the peak frequency 2106 cm‘1 with a linewidth 3.5 cm'l, and Tushaus et a1. [7] have reported an intensity z 5 I at the peak frequency 2104 cm'1 with a linewidth 4.1 cm'l. The small linewidth is also a good evidence of a clean surface. The measured peak intensity and linewidth of the C-Pt band are 0.13% and 8.8 cm'1 at the peak frequency 466 cm'1 , at the saturation coverage. Due to poor noise-to-signal ratio, the spectrum has been smoothed and a third order polynomial baseline was subtracted. The resolution was then 5 cm'l. These spectra were measured before we installed the Inductosyn, and the noise came mainly from the nonreproducibility of grating angles. These values are quite similar to those of Malik and Trenary [16]; the peak intensity was ~0.l I at the frequency 466 cm'1 and the linewith was 6 - 8 em”, but Ryberg [17,18] has reported a much narrower linewidth ~2.5 cm'1 at the frequency 463 cm'l, at the coverage 0.5, claiming that he has minimized inhomogeneous broadening by careful sample preparation. These measurements show that our system - though not yet fully optimized for very low frequencies - is capable of measuring adsorbate-substrate modes at submonolayer coverage. 5 .3 Measurements of 0-0 stretching band of 02 on Pt(l 1 1) Oxygen on Pt(111) is an excellent model system for the study of dissociative chemisorption. Oxygen is adsorbed molecularly below 100 K and is adsorbed atomically above 150 K. The molecular state that we have measured is a metastable precursor to dissociation upon heating above 150 K, leaving atomic oxygen on the surface [19,20,22] . The 0-0 bond of the chemisorbed 0; molecule is oriented parallel to the surface and is 161 greatly 'weakened from. the gas ‘phase; the 0-0 stretch frequency is decreased from 1556 cm'1 to 875 cm'1 [19] . Even though the 0-0 bond lies parallel to the surface, this mode has a relatively large dynamic dipole moment perpendicular to the surface, indicating that there is a large charge transfer between the metal and the molecule as the bond is stretched. This mode is, however, a lot weaker than the C-0 stretch mode, especially because of its large linewidth. Steininger et a1. [19] have reported a (3/2 x 3/2)R15° overlayer LEED pattern for 02 molecules on the Pt(lll) surface at 100 K. We could barely observe the LEED pattern, but it is too diffuse to tell the structure and disappears in 10 - 20 seconds from the LEED screen, probably because the structure is destroyed by the electron beam from the LEED filament. After annealing to 300 K, a sharp (2 x 2) LEED pattern is observed, which is formed by atomic oxygen adsorbates on the surface. A sharp (2 x 2) LEED pattern is also observed after dosing the sample with 1.0 L of oxygen at room temperature. At 90 K, saturation coverage is achieved after an exposure to about 0.75 L of oxygen through the doser. In the TPD spectrum two desorption peaks appear as shown in Figure 5.3.1; one sharp and strong peak at 150 K and a rather broad peak with its maximum at ~800 K. The first desorption peak is from molecularly adsorbed oxygen and the second desorption results from atomic oxygen“ This has been verified by other groups [20,21] using isotopic mixing experiments. After dosing the sample with a mixture of 1‘02 and 1802, complete isotopic mixing (160130) has been observed in TPD for the high temperature state while no isotopic mixing has been found for the low temperature state. We have observed a CO peak picking up and the 02 162 peak decreasing in the TPD spectrum if we keep the sample at 100 K even for 5 - 10 minutes after dosing the sample with 0.75 L of oxygen. Oxygen molecules stick to the surface so weakly that even at 100 K they are replaced by carbon monoxide in the UHV background. This process, however, does not occur at an observable rate if the sample is kept at 90 K for hours. Figure 5.3.2 shows IRAS spectra for several oxygen exposures. The spectra have been measured at 90 K after dosing oxygen at 90 K. The 040 stretching band is observed at 858 cm'1 at lowest coverage, but shifts upward to 875 cm‘1 with increasing coverage. The half-width of the bands is approximately 20 cm'1 and shows little dependence on the coverage. The sharp spikes are due to angular errors, and have now been eliminated by use of the Inductosyn transducer. Persson [23] has argued.that the 0-0 vibration is likely to be damped via excitation of electron-hole pairs near the Fermi level in the metal, showing that the asymmetric line shape predicted by Langreth [25] (see below) for electron-hole damping is reproduced quite well by fitting to the experimental data of IRAS by Canning and Chesters [24], and of EELS by Steininger and coworkers [19]. The conclusion also seems quite plausible considering that there is a strong charge coupling between the metal and the 0; molecule. The IRAS measurements were, however, performed using a polycrystalline ribbon sample which also lacked good surface characterization, while the resolution of EELS (50 cm“) is too large to examine the linewidth. In particular, Persson excludes the possibility of inhomogeneous broadening based on indirect and inconclusive experimental evidence. 163 Langreth [25] has derived the vibrational line shape of an adsorbate on a metal surface due to electron-hole damping and has shown that it has to be asymmetric with a tail on either side of the resonance frequency. It is assumed that the adsorbate has an induced resonance level close to the Fermi level so that electrons move in and out of the resonance level as the adsorbate vibrates and the position of the resonance level moves up and down accordingly. This leads to a breakdown of adiabaticity on the order of wr, where w is the vibrational frequency and r is the tunneling time of an electron in the resonance level. Tunneling of adsorbate electrons and holes into the substrate, thus creating electron-hole pairs in the substrate, causes the damping of the adsorbate vibration. The dynamic dipole of the adsorbate has two components; p - p*'+ u', where p1 is the contribution of ion core and.uP the valence electron contribution, and u' should have an imaginary part so that p - pl + p2 - p1(l + iwr). The line shape is expressed as: Zulzwrmu-wzrz) - 2wr(w2-w,2) IM(U) "" a (5'3'1) (wz - 0,2)2 + (0212 where w, is the resonance frequency, 1 the vibrational linewidth. The equation 5.3.1 is a generalization of the well-known Fano line shape. The asymmetry in the line shape is determined by the parameter wr, and the tail extends to either side of the resonance frequency w, depending on the sign of wr. The asymmetry parameter wr can be expressed as [25-27]: «or - (tp.u)(p‘/p1' + l)’1 , (5.3.2) where p. is the number of electrons that move in and out of the resonance level per cycle and 111' is the real part of u‘. The tail therefore extends toward the high frequency side if p‘ and pf have the same sign or they 164 have different sign with [pi] < [”1“] , otherwise it extends toward the low frequency side. Although the equation 5.3.1 has been derived for vibrational modes of a single adsorbate perpendicular to the surface, the same equation applies to a layer of dipole-coupled adsorbates with renormalized parameters [26] , and to vibrational modes parallel to the surface [27]. Persson [23] has derived the same Fano line shape for the electron-hole damping of adsorbate vibrations on metal surfaces using a somewhat different approach. The isotope dependence of the linewidth 1 dominated by electron-hole pair damping has been predicted to be characterized by a - -1. Here a is defined as a - (aln 1/aln m,), where m, is the reduced mass of adsorbate (see section 1.2). The linewidth 1 for electron-hole pair damping is expressed as [25-27]: 1 - 21rm,.(6¢p.)z , (5.3.3) where 66 is change in the energy of the induced resonance level of the adsorbate caused by stretching of the bond, and p. is the number of electrons that move in and out of the resonance level per cycle. Since 56 ~ Q0 ~ (m,w,)1/2, the linewidth is inversely proportional to the reduced mass: 1 ~ mt'l. As discussed in section 1.3, the mass dependence is probably the best way to sort out the inhomogeneous broadening. We are therefore trying to measure the isotopic effect on the linewidth, in order to further understand the line broadening mechanism. We expect to see that the linewidth for “02 is smaller by 1.2 cm'1 than that for 1‘02 if it is dominated by inhomogeneous broadening, or by 2.5 cm'1 if it is dominated 165 by electron-hole pair damping. In Figure 5.3.3, the solid curve shows the best fit, using the equation 5.3.1, to the measured IRAS spectrum for M0; on Pt(111) at the saturation coverage N'- 6 x 101‘035 coboefi 0mm cor cm? 00.. can OOH omw com om; on: a: m A q _ q a a an. _ d E e a fi fi .. a a E r. e a .. . t t a t e a r. e z a p D. 3 N \ a m K’s/>712 N s s .. a. 00 U nu... 0 fix 0 one $0.21??— L L. _ fr _ _ _ L _ 172 r- a A r ‘ E 3. ,- i E , . 3 i s t i U) 8 r . r 1 ' 1 WW! I a m J; J A _L a a L 0.1 0.3 0.5 0.7 0.9 1.1 Temperature (K) Figure 5.2.1 A TPD spectrum measured after dosing the Pt(lll) sample with 0.5 L of CO gas using the doser. 173 0.005 L 0.02 L 0.1L 0.5 L 2 D wlw ' I ' I v v . %%fi 4 i L A 1 r 1 2050 2080 2110 2140 2170 2200 Frequency (cm-1) Figure 5.2.2 Spectra for GO stretch mode of on-top CO on Pt(111). CO gas exposures are in Langmuir and are referred to ion gauge readings. The effective exposures at the surface were larger by a factor 20 because of the doser. A linear base line is subtracted for each spectrum . 690833 mm on: 83 a can 858.5. mm 8.58% 25. .286 9.282% 560 05 .«o 8.58% mg m.~.m Semi A 2-83 honoeuoum 174 m _m E? mt. «we _mv c _v _ 4 _ _ . _ . J L WflGOAUI 1 :86- 1 Seed. W - mocod- .. \2/ / \itt \/\./ 1 886 L » F F _ . _ . p . mocc.c 175 r- -< r- a .g t- q '3 r- 'i 3 ‘2 T '“ .9 m g i‘ X10 1 r- 4 [- ‘1 2 L l 0.2 0.4 0.6 0.6 513.0 Temperature (K) x Figure 5.3.1 A TPD spectrum measured after dosing the sample with 0.8 L of 02 gas. The sharp strong peak at 160 K is from molecularly adsorbed oxygen and the rather broad peak at about 800 K is from atomic oxygen on the surface. 176 0 l 8 030L - a ‘9. r o .3 osor , < 0.701. 025% 810 840 870 900 930 960 Frequency (cm-1) Figure 5.3.2 IRAS spectra of the O-O stretchung band for different 02 exposures. Spikes are caused mainly by the nonreproducibility of the grating angles (see also section 4.7) 177 .8588— ocam 05 o. E v.8 .8838 85889. 85 8 8888 22?. :0 89.88 88888 NOE 8.5 .8589? 5.2% 0.0 05 no 858% mg 5. m.m.m 2am."— A.-.Eov 88:70.."— omc _ com sea 3% cf. can . i . . _ 1 _ . _ . $25 a. \ r nnccd. \ f A r c256. a. . 5.: _ .1386. W .J .45. "A‘ .a :4 fit... 1 c—ocd. L _ _ _ e _ r L » “cocoa 6. Conclusion Many interesting surface vibrational experiments require high resolution, high signal-to-noise performance and excellent baseline subtraction, especially in the low frequency region. Instruments with those capabilities are not readily available. We have thus built a new far- infrared reflection- absorption spectroscopy system designed specifically for the frequency region 350 - 1000 cm”, though it can be used up to 3500 cm'l. It incorporates a liquid-nitrogen-cooled grating monochromator and other cooled optics, a Si:B photoconductive detector and a UHV chamber equipped with LEED, Auger and TPD. We have measured the 0-0 stretch mode of molecularly adsorbed 02 on Pt(111) as a function of coverage and the molecule-substrate mode of CO on Pt(111), showing that the system is working well. 3 The noise-to-signal (N/S) ratio achieved to date is 2 x 10" l/fl-E and the dominant noise source is the nonreproducibility from one scan to the next. An improvement of the N/S ratio by a factor of 4 is expected by improving the scan-to-scan reproducibility. As a first experiment, the isotope effect on the linewidth of the O-O stretch mode will be measured. The shifts in the linewidth are expected to be 1.2 cm'1 if inhomogeneous broadening is dominant, or 2.5 cm'1 if electron-hole pair broadening is dominant. With the resolution, N/S ratio and baseline subtraction performance of our new system, those shifts can be measured. A broadband infrared absorption measurement is planned as the next experiment. The broadband infrared absorption is known to have a 178 179 characteristic dependence on. the frequency [for' metals, and. to have correlations with surface resistivity, depending on whether there are adsorbates or not [1]. In addition, we will perform measurements of adsorbate-substrate vibrations of atomic and.molecular adsorbates, alkali metal adsorption on metal surfaces, and adsorption on semiconductors, which require high resolution, very high signal-to-noise ratio and excellent baseline subtraction. 180 Reference 1. K.C. Lin, R.G. Tobin, P. Dumas, C.J. Hirschmugl and G.P. Williams, Phys. Rev. B, to be published, and references therein. "00000000000