I???“ .33: ”4‘“!qu I I "4 '-' -. J's}31" '1 .9 .4 I: he: I ' *l'i‘IIIIIl34fi-Iu‘mi3rg, 3 2:; ‘ ‘ ‘ . : :J .'.' . .“hfi _‘_.‘. , 2:30}: :I7_‘;,\.,' ‘2' .34.: 7 . am ,A_ -N .r__ .,':.~1§ ‘:hII: : 3" 535; '2’: 41 .4 .. ’4‘: 4. “3a.; Fig-II flzr'gflflge 83,133? ‘43“:4433433 r, . 43,, , .93: ' 1. . < :3 4I ~ "1 n “N L ' $3, firfig‘gn 31f x1331} ' I‘ I {I‘ ~ 4:15 ~ $44454 . '1 Strl‘gznxn ; ‘.....r ‘ ‘ '1’.‘ 34,5 . 8 , 533‘"! 3:25." I? .23. 1 Y‘.‘ I q r“ . ' q ‘1 3;. v4, __.‘ = h p- p 5‘. ' N ‘—- 12‘... ‘- __>‘-—-.~ . . . .4». ‘..... «1 ’ u. . "'..—‘..: ‘.." ‘4 __. - d... _ - ~v-,_u ‘ _~.—~— -1—‘v- —- . .~_ ”'..:— ‘ ow » . _:I7 . ... 37—93}, L. 0.5....“ —I ’5...— :M.‘ . -H-i ,- “4.— _1 _ ~.....- . v .. - c a» I}, ' ,Ie. IIII 33:“ I'I III‘III IIIIIII :33I:3:336I: 333‘” 3:: I “PM: .' “t I :9 M-‘II: :I'I.I 'I ~ 3 :::'Is=:=:::444IN II” I II ':"I 334:. 2‘ ~r ‘---.. v.4” .-.._, p 0‘ nd. This arrangement is known as the Kretschmann34 configuration. When electro- magnetic radiation traverses the prism at the correct angle ¢ (according to Snell's laws) it undergoes total reflection at the interface with the dielectric medium. An evanescent wave is generated at the prism-dielectric interface by virtue of conservation of momentum constraints. The evanescent wave propagates through the dielectric spacer layer with a phase velocity C/np. Since np is large the low phase velocity of the evanescent wave converts it into a radiative surface plasmon at the metal-dielectric inter- face (for low k values). The radiative plasmon may then couple to electronic transitions in the molecules of the spacer layer. Several reviews of surface plasmon coupling have appeared in the literature (e.g. see references (25) and (35)). Experimental verifications of enhanced Raman scattering from adsorbates through ATR coupling of surface plasmons abound in recent publications.36-44 12 Though coupling to surface plasmons has been proven to enhance the Raman scattering intensities, the enhancement 4--106 reported by factors observed fall far short of the 10 van Duyne. Indeed, even exactly on resonance surface plasmon coupling can only account for a factor of 10 to 100.38 Ultimately, the giant Raman intensities observed may be accounted for by several cooperative mechanisms. Other theories to explain the anomalous surface Raman intensities have been proposed which have not enjoyed the widespread acceptance of either the image field theory or surface plasmon coupling. These include resonance Raman 45,46 from surface transients or radicals, normal Raman scattering from molecules intercalated in a dense carbon matrix at the electrode surface,47-49 resonance Raman scattering from slightly perturbed surface and molecular states,50 and radiative excitation and recombination of 51'52 Of these, Cooney's work on the particle-hole pairs. role of surface carbon helped researchers recognize the important influencecfifsurface impurities and correctly cautioned coverage interpretations. Regis and Corset46 have observed enhanced surface Raman scattering from bipyridines on silver. Their assignments were particularly useful in the assignment of the Raman spectrum of t—l,2- bis(4-pyridyl)ethylene on silver in this report. Studies of adsorbates deposited under ultrahigh vacuum conditions have provided considerable information on 53,54 coverage dependencies of intensities, orientation of 13 molecules at the surface,55 energy transfer between adsor- bates and meta1,56-58 influences of impurities,47 roughness phenomena,59 and luminescent background identification.60'61 Demuth et a1.55 examined an interesting coverage dependent phase transition on silver. When the coverage was increased the Raman spectrum was interpreted to indicate a transition from flat molecules to end-on configured pyridines. Con- flicting reports from Bell Laboratories leave the question of distance for maximum intensity enhancement from the metal surface unresolved. Smardzewski et a1.53 initially reported that molecules closest to the surface exhibited the largest intensity enhancement. This conclusion was reached by observation of Raman spectra from a monolayer of deuterated pyridine followed by increasing coverages of undeuterated pyridine. The deuterated pyridine closest to the surface exhibited the most intense Raman scattering. Rowe et a1.58 measured the enhancement as a function of molecule-surface separation reaching the opposite conclu- sion i.e. molecules beyond those immediately adjacent to the metal surface also exhibited Raman intensity enhance- ment. The controversy is representative of the overall conflict currently in the SERS field. That is, is the enhancement effect a chemical one, requiring the formation of a chemical bond to the surface and enhancing only those molecules attached to metal atoms? Or, is the enhancement factor a result of electromagnetic phenomena, in which the field at the molecular position is enlarged due to the presen in the centre direct CN-, a other Pyraz contr surfa pyrit rule 14 presence of the surface, thus, allowing molecules simply in the vicinity of the metal to become enhanced? This controversy remains unresolved. Although the greatest concentration of effort has been directed at the study of pyridine, pyridine derivatives and CN-, applications of surface enhanced Raman scattering to other molecules have also appeared in the literature. 62'63 for example, has provoked considerable Pyrazine, controversy over the appearance of ungerade modes in the surface Raman spectrum. Similar to t-1,2-bis(4- pyridyl)ethylene, pyrazine has a center of symmetry and the rule of mutual exclusion is operative. Forbidden bands are also observed in the surface Raman spectrum of benzene which have been discussed in terms of symmetry lowering 64'65 Biologically important molecules have been observed on silver66-68 and resulting from surface site symmetry. demonstrate advantages of surface studies. Some benefits of surface studies of biological molecules are the absence of flourescent background owing to the non-radiative excited state decay pathway provided by the metal substrate, and general absence of decomposition problems from over- heating in the beam with the metal surface acting as a heat sink. Moskovits and DiLella69 observed surface Raman scattering from ethylene and propylene on silver. These spectra were helpfulithhe assignment of t-1,2-bis(4- pyridyl)ethylene SERS. Several other molecules including water,70 €0,71'72 and EDTA73 have also been observed on 15 silver substrates. Several of these enhancement mechanisms and surface 74,75 techniques have been reviewed in the literature. In this thesis only the mechanism of resonance Raman enhance- ment by surface complexes is addressed. CHAPTER III THEORY OF RESONANCE RAMAN SCATTERING FROM SURFACE COMPLEXES. One mechanism which has recently advanced to the forefront of surface Raman discussions is resonance Raman scattering from surface complexes. The formation of a medium strength chemical bond between cyanide and the 76’77 However, silver surface has been demonstrated. similar efforts to observe a silver-pyridine stretching vibration have generally failed. In this chapter the implications of surface complex formation on the intensi- ties of Raman signals from adsorbates is discussed. Complex formation at the metal surface requires locali- zation of metal electron densities. Such localization was examined by Goddard and McGill78 in a review of quantum chemical methods appliedixasurfaces. The model found suitable for surface electron localization was intermediate between a lone metal atom and the delocalized metallic matrix, i.e. a cluster of metal atoms. With the orbital localization suggested by cluster formation a connection is made between the roughness requirements of surface 79-81 ' 82-84 enhanced Raman scattering, the adatom hypotheses, and a surface complex induced resonance Raman enhancement l6 17 85'86 That is, clusters of metal atoms formed in mechanism. surface preparation procedures (e.g. oxidation-reduction cycle) result in orbital localization which may be utilized in overlaps of adsorbate electron densities. Metal-adsorbate electron density overlaps may signi- ficantly alter the distribution of electronic states. Thus the possibility of coming into resonance with a low lying excited state of the complex even with red excitation becomes very real. Therefore, mechanisms of resonance Raman scattering warrant further study. According to 87 Rousseau et al. the total scattering cross section is 8nw 4 2 _ s 2 l p K 0 — 4 fi- (5 _w ”if ) + ar (1) 9cl1 e,v g,e L I where the bracket represents the pKth component of the molecular polarizability, apK; wL is the laser excitation frequency; ms is the scattering frequency. gi (electronic state 9 and vibrational state i), ev and gj represent respectively ground, intermediate and final vibronic states in the scattering event. re and rK are electronic position operators, with p and K representing incident and scattered polarizations respectively. PI is the linewidth of the intermediate state and ar represents antiresonance sums. The sum runs over all vibrational-electronic states except initial and final states, i.e., Igj> + |gi>. When the frequency of the incident photon matches the frequency 18 of the electronic transition (resonance condition) the scattering cross section becomes large and intensities increase. Most of the adsorbates examined in SERS experi- ments exhibit no resonance Raman scattering in solution owing to the lack of low lying (visible) electronic states. Therefore complex formation must be presumed if the enhance- ment mechanism of resonance Raman scattering is invoked. The metals which have demonstrated surface enhanced Raman 10 metals. This scattering (Cu, Ag, Au, Hg) are all d wealth of electrons may be used to occupy normally unoccu- pied adsorbate molecular orbitals upon complex formation as observed for the Ru(NH3)5(t-l,2-bis(4-pyridyl)ethylene) complex88 (see Figure l) in which n* states in the bipyri- dine ligand are occupied. In the pentaminebispyridylethy- lene ruthenium II complex the complex symmetry is CS removing degeneracies occurring in the C2v pentamine ruthenium substrate. The attachment of t-l,2-bis(4- pyridyl)ethylene to the silver surface should similarly reduce the symmetry of the molecule probed in the Raman experiment. Perturbation of the potential energy of the normal modes in the adsorbate upon complexation may also be invoked to account for minor shifts of vibrational frequencies observed in the surface scattering experiments. When considering the electronic distributions of states in a surface complex in more detail some important conclusions can be made. Excluding for the moment the influence of the image dipole, formed by rearrangement Figure l: 19 Energy State Perturbation in t-l,2-bis(4- pyridyl)ethylene upon Complexation to RuII (NH The point groups for the ligand, 3’5“ complex and Ru (NH3)5 are given. Ru(II) has a d6 electron configuration and thus the lowest three levels in the energy level diagram for the complex are occupied. 20 Azzevnaet NI a.w$-u (3) and the corresponding moment of K polarization in (3). (mp(Q))g e is the pure electronic transition moment I (mpmng,e = <9 r le > (4) at nuclear configuration Q. In order to integrate Equa- tion (3) over nuclear coordinates the explicit form of the Q dependence in (mp(Q))g e is required. Following Albrl seril Subs1 into In a the e Equat theor Eqn. Rayle 27 Albrecht92 and Rousseau87 (mp(Q))g e is expanded in Taylor I series about the nuclear coordinates amp Q (m (0)) = (m (Q l) + (———) K (5) o g:e p 0 g,e 30k O,g,e, Substituting the result into the transition moment and into Equation (1), the molecular polarizability becomes 1 <1jv> a = E (m (0 )) (m (0 )) — 2 _ _ pK e p 0 g,e K 0 g,e h v (mg,e wL Fe) + 2 (ame) (amK) l E + e BQK 0 BQK 0 h K (wg'e-wL-ire) (6) In a normal Raman experiment in which the v-dependence of the energy denominators is negligible the integrals in Equation (6) may be evaluated by invoking the closure theorem (i.e. 3|v> = % lv> the first term in I Equation (6) becomes 1 we|e>|v> A' = 2 (mp(Q0))g'e(mK(Q0)) 2: e759 g’e H v (8) The Hamiltonian in Equation (8) can be expanded about the equilibrium nuclear configuration following Clark and Stewart.94 %- =%- + (335900 + — weo e = % + 00 + ...-weo and =11? oQ (9) * Note: the damping factor has been omitted since the expansion is strictly only valid far from resonance. 29 since % = meo. This leads to a revised A term which includes the resonance Raman case: l O A'=E(m(Q)) (m(Q)) 2- p 0 g,e K 0 g,e h 0_ _ 2 (10) Thus, if a force, given by the derivative of the potential part of the Hamiltonian* with nuclear coordinate (Fe0 = (3Ve/BQ)O), is experienced upon excitation into an excited state, then the first term in Equation (6) may contribute to observed Raman intensities. To experience a force in the excited state the equilibrium position in that state must be shifted relative to the ground state equilibrium position. The shift in equilibrium position will be only along totally symmetric normal coordinates since the matrix element vanishes for non-totally symmetric coordinates. Thus, resonance Raman scattering by the A' term will result in enhancement of totally symmetric vibrations. In the second term of Equation (6), Albrecht's B term, the dependence of the electronic wavefunction on nuclear coordinate is explicitly contained in the coefficients . 92 (amp/BQk)0 and (amK/BQk)o. Follow1ng Albrecht the * Note: In the adiabatic Born-Oppenheimer approximation the potential energy of a nuclear vibration is equal to the total electronic energy therefore (aVe/aQ) 0 = o. 30 coordinate dependence of the electronic moment may be treated as a Herzberg-Teller perturbation i.e. 8m (4343”2 = >3 mg'smoi ,1; . (11) 3 30k Therefore, it is evident that contributions to the Raman intensity from both the A term and the B term depend on the quantity (aflyan). In the A term (axyao) produces vibronic mixing within one excited state and in the B term (aflyan) couples different electronic excited states. Since the overlap of energy bands in the electronic manifold of the valence and conduction electrons is extensive in the metal and limited in the adsorbate the magnitudes of (Bx/an) in the adsorbate and image dipole are likely to be quite different. Since (BHYSQ) is responsible for the intensity of the resonance Raman scattered light that intensity emanating from either side of the silver surface will not be the same. Therefore, contrary to the prediction of the "surface selection rule", the image field can only act to reduce the symmetry of the scattering system not increase that symmetry. Thus, with the absence of any symmetry ordering influence by the conduction electron image, the system being probed in the Raman experiment returns to a one- sided attachment of an adsorbate to the surface. Symmetry reductions caused by the one-sided nature of the system, 31 the mismatch of the image field, and low site symmetry experienced by an adsorbate at the surface of a metallic matrix, result in expectations of extremely low symmetry scattering systems. The most common point group expected, even for highly symmetrical adsorbates, is C1. Of course, in such low symmetry all modes will become Raman active and, in general, an increase in the number of observed Raman lines is expected. However, in a resonance Raman mechanism only those modes effective in coupling electronic excited states through electron density overlap are enhanced. Thus, all possible vibrations may still not be observed. For example, t-l,2-bis(4-pyridyl)ethylene adsorbed at a silver surface may actually be attached through either of the nitrogen atoms (but not both) in an "end-on" configuration or may lie flat on the surface. In the end on configuration only the nitrogen atom experiences electron density overlap with the metal. Thus, it might be expected that the surface Raman spectrum observed would not be much perturbed from the normal Raman spectrum. In the flat configuration, in contrast, the entire molecule experiences some overlap of electron density with the metal and all vibrations (both infrared and Raman active) are expected to be observed. Conclusions drawn in this chapter are extrapolated to the discussion of orientation and enhancement mechanism of t-1,2-bis(4-pyridyl)ethylene at a silver electrode surface. The experimental surface Raman spectra and discussions are presented in Chapter 6. 32 The surface complex enhancement mechanism has been slow to attain fruition. This has mainly been the result of a lack of success in observation of a metal-adsorbate stretch in the low frequency region. For all surface Raman studies a very complicated low frequency region is observed which has hindered assignments. The appearance of a strong, Ag-Cl stretching mode in electrolyte studies near 230 0111-1 also may cloud assignments in the region. Pettinger and Wetzel85 recently applied reduced potential difference spectra (RPDS) to the low frequency Raman spectrum of pyridine on silver. In doing so the Bose- w/kT - l).1 in the prefactor of the Raman Einstein term (e cross section could be eliminated revealing hidden struc- ture in the low frequency region. A silver-pyridine sym- metric stretch was assigned to a 210 cm.1 line. Efforts to find a metal-adsorbate stretch may be doomed at the outset. Such a mode can be expected to be considerably damped owing to one atom of the vibrating pair being associated in a large metal matrix (see, for example the treatment of free damped motion by Rossgs). The low amplitude resulting from damping will result in low Raman intensity. Forces associated with the electric field at a metal electrode also may act to damp the oscil- lations of an adsorbate metal bond. Experimental evidence of surface complex formation in the system t-1,2-bis(4-pyridyl)ethylene at a silver 33 electrode is presented in this thesis. Assignments are made without regard to the "surface selection rule". CHAPTER IV Experimental Raman Spectrometer All Raman spectra reported in this thesis were obtained with a Spex Industries model 1401 Ramalog system. Photon detection was accomplished with a cryostatted (-20°C) RCA model C31034 GaAs photomultiplier tube in photon counting mode. All spectra were recorded with a resolution better than 2 cm-l. Samples Polycrystalline samples and aqueous solutions were sealed in 1.7 mm capillary tubes. High vapor pressure methylene chloride solutions were placed in a Raman cuvette. A translating Raman cuvette holder and brass cooling jacket were constructed for optimal sample management. Opaque polycrystalline samples were observed in a back scattering geometry. The incident laser beam was allowed to pass through a hole in a 45° mirror just prior to striking the capillaried sample. The back scattered radiation was thus reflected into the entrance slit of the double monochromator. 34 35 The first electrochemical cell utilized in surface Raman studies was constructed from a hollowed-out lucite cylinder. Raman spectra were observed through a quartz optical flat. Because of difficulties in maintaining the system free from impurities, the lucite cell was replaced with a glass cell which could be cleaned by immersion in concentrated nitric acid solutions. The working electrode was a polycrystalline silver disk fitted to a teflon sheath. The surface area of the silver disk in contact with electrolyte solution was measured to be 0.11 cmz. The working electrode was generally placed very close to the observation window only after the oxidation- reduction cycle was complete. This was done to prevent non-uniform coverage of the electrode surface caused by concentration gradients which develop during the cycling procedure. Subsequent movement of the electrode surface close to the observation window permitted observation of maximum intensity from surface species and a minimum from the bulk solution. The reference electrode was a silver-silver chloride electrode. This was prepared by electrolyzing a coating of silver chloride onto a silver coated platinum wire which was immersed in a saturated silver chloride solution (prepared by addition of a few drops of silver nitrate to a saturated KCl solution). The body of the reference electrode was k" glass tubing to which an ultrafine mesh porous plug was attached. Generally, the reference electrode was 36 attached to another k" glass tube containing the supporting electrolyte. This prevented potential drift accompanying minor leakage. The reference potential of the Ag/AgCl couple was 50 mV relative to the saturated calomel elec- trode. All electrode potentials reported in this thesis are referenced to the saturated calomel electrode (SCE). In supporting electrolytes containing NaClO4 the reference electrode was prepared with saturated NaCl solution instead of KCl so as to prevent precipitation of KClO4 which is minimally soluble in water. The counter electrode was a heavy gauge platinum wire. Nitrogen gas (Airco Inc., 99.99% pure) was bubbled into solutions prior to electroly- sis and often during Raman observation to limit the forma- tion of silver oxides. All surface spectra were recorded at an angle of incidence of 60° for maximum scattering intensity as observed by Pettinger et al.97 This system is shown in Fig. 3. The electrochemical cell was controlled with an Eco Control Inc. model 550 potentiostat and model 731 digital integrator. Generally, 50 mC/cm2 (or 5.5 mC for this electrode) were allowed to pass during the oxidation- reduction cycle (i.e. the potential was maintained at +100 mV versus SCE until sufficient charge passed and then the potential was returned to -600 mV versus SCE). This cycling yielded a rough electrode surface. The reflecti- vity of the metal surface drops during the oxidation- reduction cycle producing a slightly yellow coating even 37 in the simple chloride blank. Varying amounts of residual charge remained after completion of the ORC depending on adsorbate and electrolyte. Potential drift was negligible (~ct2 mV) over the course of a typical Raman experiment. Potential cycling was carried out in the presence of adsorbates. For solutions of t-l,2-bis(4-pyridyl)ethylene at millimolar concentrations potential cycling produced a dark brownish-black coating at the electrode surface. However, similar Raman intensities were observed for 500 nM solutions of his pyridyl ethylene which upon cycling yielded a yellow surface coating undistinguishable from that observed in the chloride blank. The silver electrode surface was subjected to mechani- cal polishing before each experiment. This consisted of applications of A1203 powders (Buehler Ltd.) of diminish- ing size finishing with 3 micron powder. The electrode was then rinsed with methanol and water and transferred to the electrochemical cell with a drop of water attached to inhibit oxide formation. Lasers A Spectra Physics model 165-08 argon ion laser (4 watts all lines) with model 265 exciter was used for 5145 A excitation and to pump a continuously tunable Spectra Physics model 365-09 dye laser with three-plate bire- fringent filter. A Classen filter was used to assume monochromaticity of dye laser emissions. Rhodamine 6G 38 Figure 3: Experimental arrangement in the SERS experiment. R = reference electrode (Ag/AgCl); W = working electrode (polycrystalline silver); C = counter electrode (platinum wire); FL = focusing lens; CL = collection lens; ¢- = angle of incidence lnC relative to the surface normal. 4O (Exciton Chem. Co.) was the dye used in obtaining excita- tion profiles of surface molecules covering the range 5600 A to 6600 A. For red excitation a Spectra Physics model 164-01 krypton ion laser was employed. Excitation Profiles Excitation profiles of species adsorbed to the silver electrode surface were obtained in order to examine the electronic manifold of states in the surface-molecule system. The 935 cm.1 line of sodium perchlorate (0.5 M) served as an internal standard. The intensity of the 935 cm.1 line could be varied by positioning the electrode surface at various depths in the electrolyte. Because of the potential dependence of the surface concentration of C104- a better choice of internal standard is necessary for comparison of potential-dependent excitation profiles. The teflon sheath fitted to the silver electrode serves this purpose well. Both the metal surface and the teflon could be illuminated simultaneously with the application of cylindrical focusing. The Raman spectrum of teflon is reported here and is generally uninterfering with the surface Raman spectrum of t-1,2-bis(4-pyridyl)ethylene (see Figure 4). It became necessary in the course of carrying out excitation profile experiments on surface adsorbates to assure that the laser beam remained focused to the same point on the surface with each excitation wavelength. 41 Figure 4: Normal Raman spectrum of teflon and surface Raman spectrum of t-l,2-bis(4-pyridyl)ethylene on silver with teflon standard. 42 is... asses Essa 5:: ...; a...” s we... :3 8e - q q _ £49 q Om.“ - £9— . 3N.— “ fi‘u _ 8W“ ......m m .. : m z m n m n m 5:2 mm A ...... m. u E. J T. 43 Since tuning the dye laser requires passage through a birefringent filter of refractive index greater than 1 each excitation frequency will traverse space with a different vector (similar bending occurs during passage through the Classen filter). Therefore an aperture was placed just prior to the sample and the laser direction adjusted to pass through the aperture and thus illuminate the same point on the surface regardless of excitation frequency. Spectra were recorded every 100 cm.1 during the excitation profile. The excitation profiles reported in this thesis are corrected for the v-4 dependence of the Raman scattered light intensity.5 Spectroscopic intensities were norma- lized to the intensity of the 935 cm.1 line of the perchlorate, but not normalized relative to intensities at different excitation frequencies since doing so would effectively double the relative error in each profile data point. Therefore the most intense peaks in the profile belong to the most intense peaks in the spectrum itself. A program for v-4 correction and normalization was written for a Hewlett-Packard 67/97 calculator and is available upon request. Laser Power Measurement Laser powers were measured with either a Photodyne model 44XL optical power meter or a Coherent Radiation model 210 power meter. 44 Visible Absorption Spectra Absorption spectra of t-l,2-bis(4-pyridyl)ethylene, its dihydrochloride and a silver complex of t-BPE were obtained on a GCA/McPherson model Eu-700 series UV/Vis spectrometer. Sample Preparation: t-l,2-bis(4-pyrigyl)ethylene(t-BPE) Trans-1,2-bis(4—pyridyl)ethylene (Aldrich Chem. Co.) was recrystallized from hot water; some orange color was removed in the process. Normal Raman spectra of polycry- stalline t-BPE were recorded. Since t-BPE is only sparingly soluble in water, reaching saturation at about 3 mM concentrations, spectra were recorded in methylene chloride solutions. Solutions for surface studies ranged from 2.5 MM t-BPE (heavy coverage) to 250 nM t-BPE (light coverage). Supporting electrolytes were either 0.1 M KCl or 0.1 M NaCl and 0.5 M NaClO For "light coverage" 4. surface studies fresh polishing pads were used owing to the presence of residual t-BPE on used polishing pads, which could influence bulk concentrations and therefore coverage. t-l,2-bis(4:pyridyl)ethylene dihydrochloride and dideutero- chloride Trans-1,2-bis(4-pyridyl)ethylene dihydrochloride was prepared by dissolving t-BPE in methanol and adding concen- trated hydrochloric acid until no further crystallization was evident. The pinkish orange crystals were collected on 45 a 15 mesh buchner funnel and washed with methanol. The Raman spectra of polycrystalline and aqueous solutions of the dihydrochloride were obtained. Trans-1,2-bis(4-pyridyl)ethylene dideuterochloride was prepared by bubbling DCl into a methanol solution of t-BPE. DCl was prepared98 by slow addition of 5 ml D20 to 30 ml benzoyl chloride and refluxing under light heating. Unreacted D20 and benzoyl chloride were collected in an ice trap and DCl in a liquid nitrogen trap. DCl was bubbled through the t-BPE solution until no further crystallization was evident. The pinkish-orange crystals were collected and washed with methanol. A similar prepara- tion was attempted by using chloroform as the solvent and a pink compound was obtained after considerable DCl was passed through solution. This compound was insoluble in water and was air oxidized overnight to a yellow crystal- line compound. It is believed that this pink compound was the addition product of DCl across the ethylene double bond. However physical evidence of its structure is currently incomplete. Silver Complex of t-l,2-bis(4-pyridyl)ethylene (Agn(t-BPE)m_ A silver complex of his pyridyl ethylene was prepared by addition of an aqueous solution of silver nitrate to an ethanol solution of t-BPE until no further crystallization occurred. Crystallization appeared complete after addition of silver nitrate to 1:2 (AgNO3/t-BPE) molar ratio with 46 t-BPE. Crystals were collected and washed with both ethanol and water. The complex was found to be soluble only in n-butyl amine and pyridine. The silver complex was found to be insoluble in water, ethanol, CC14, CH3C1, CHZCl, CHC13, dimethyl sulfoxide, tetrahydrofuran, ammonium hydroxide, methyl amine, n-propyl amine, C82, methanol, acetone, cyclohexane, n-hexane, diethyl amine,and benzene. The Raman spectrum of the polycrystalline silver complex was obtained and is reported here. Amino Acid Solutions Amino acid solutions were prepared to 0.05 M'in 0.1 M KCl stock solution. Observation of Raman spectra of the amino acids on silver adsorbed from neutral solution was unsuccessful. However spectra were observed from solutions in which the pH was adjusted above the isoelectric point of the particular amino acids. Glycine and leucine surface spectra were recorded at a bulk solution pH of 11.0. Agonz Agzg Raman spectra of polycrystalline silver hydroxide (AgOH) and silver oxide (A920) were obtained for comparison with surface spectra to examine the possible occurrence of surface oxide formation. Commercially available silver oxide (Fisher Chemical) was used without further purifica- tion. Silver hydroxide was prepared by addition of excess hydroxide (KOH) to an aqueous solution of silver nitrate until the solubility product of silver hydroxide 47 (1.52 x 10-8; 20°C)99 was exceeded. Both silver oxide and silver hydroxide are black polycrystalline substances and therefore a back scattering geometry was required to obtain the Raman spectrum. The specific geometry was described previously in this chapter. CHAPTER V Polycrystalline and Solution Raman Spectra of t-l,2-bis(4-pyridyl)ethylene, its dihydrochloride and its dideuterochloride In order to understand the surface Raman spectrum of trans—1,2-bis(4-pyridyl)ethylene (t-BPE) more clearly the normal Raman spectrum was recorded. The Raman spectra of the dihydrochloride and dideuterochloride are also reported to assist in assignment of vibrational modes. The Raman spectra of these compounds have not previously been reported in the literature although the infrared spectra 1’100 and cisl isomers have been reported. of both the trans Both solution and polycrystalline spectra are reported. Solution depolarization ratios are also determined. The Raman spectra are shown in Figures 5 and 6 and the vibra- tional frequencies are listed in Table 1. The pyridine ring vibrations are assigned according to Wilson notation.101 All normal Raman spectra reported here were obtained with 6471 A excitation. Symmetry Considerations The molecular point group symmetry of t-l,2-bis(4- pyridyl)ethylene is C2h' The center of symmetry in the molecule requires that infrared and Raman activities be 48 49 mom moo moo mam Hem mum Hmm mvm mow mam Ohm «om mwm vmm HVH 00H mm on mm vm mv H OHNH H H mom Ohm mom Nmm Hmm ovm Nov mmm UmHOm O N N HO No D mam!» .mmfifiHMCOU H QHQMB momH amm. omma meoa moom amm. mmm mmm mmm emm mmm deem. mmm mme name. mme mee mee ome Hoe omm emm oem omm emm mm om me nmmom a 0mm numaomm mmmubunu mama mam. mmam .nbn e..o me me emaa Ma meom nmm 4 Ha mam Ema. mmm m we Ha emm omm nmm mm Hm mmm emm m on ma mmm nos mm ma mmm meme. omm mom mm m mme noes. mew be mm mm flee amm. one no ma Ha eme gem mm mm med n ma omm mam mmm x N am oem mmm nponnnm> mom mnbmaexm Ham Hm WOCOZ wombbnq me neaom e mommo Am nowamse awn nmo >mo iiiiiiii mmmiuiilil ucoEcmmmmm ilmuuoafiwmi .< anew "COmumumoxm .mpmuoHcooumusopmp ocoamnuoaawpmnwmivvmmb:~.aiu pcm mpmuoHsooupwnmp ocoawnuwaawpmu%mivvmmbnm.aiu .wcmawnumfiawpmuwmlvvmmbum.alu mo muuoomm cmEmm "a mam¢a 50 maom moam aeem amom emom nneamm neoeem mmea amea meea amea mmea mmea eoea mema nnmoma mama aama meea emma aama eema mmma mmma eema aama mama eaaom 0mm aoma mam-» uneoem eemm nnommm mmom aoam amom omem amem emea mmea eeea nem. meea mmea nem. meea moea anooma anaoma mama nem. mama mmma nee. mmma eema mam. aama mmma omma emma mom. oema m eaaom e o n :lmao m mamlbulu emom oeom amom emom maom emea amea mmea amm. aaea aaea omea mama nae. mama aama nem. aama aama aaea nma. maea aaea amm. mmea aaea mam. aaaa amma mmm. mmma mama aama naa. mama emma nmm. mmma mmma Ema. omma eaaom a aommo uuuuuuuu manipulnuu .nbm muz a m< A..Hu.m UHU x NV AN < noumnum mo .aun mlmo a9 m4 aooma x my .aun auz .nbm e> m Ma mm mm .uum ono < a a< am am me am he < mm e x m we am naa < a< baa we mm aa + e 4 a< aa x m w< am ea < mm be x m we a< m mm mm .mm@ mz ha mm mm ea c am 1* :onaaze nma nmo >mo ucmecmmmmfi ilmuumfifiwml .poscmucou a canoe Figure 5: 51 Normal Raman Spectra of polycrystalline t-l,2-bis(4-pyridyl)ethy1ene, its dihydrochlor- ide and dideuterochloride. + EXCITATION WAVELENGTH: 6471 A (Kr ). lllI) 9» fl e\ c: - -~ 3054 w ?3038 -3 g_ r 013 O A 1676 \ __ -—1630 16" g;- f 1591 c: , 1549 1541 * ~1491 :;’ .1319 8' K iii? { —1319 ‘ - 123: a (‘ 1225 :5. --—1198 Q t. ”80 § \rm —995 ‘958 300 ’> 878 a °.,, 0 \ ‘1-611 641 _. a: :'—- °-_ 0 «=1 ’7’ 2’. ‘31 ‘F 005 :3 :L =- ‘C 3 s o- 0-0 370 {if 0 = O 200 25 200 O f‘:) 121 -, —;:—-— ~—-="9‘ 4:: _____,3 52 “L 3005 3059 3050 jF’3031 1571 44 g ‘19 -————-1535 -1605 ..1512 1500 1373 1346 1333 ‘ - 1254 . -— 1203 1062 ——aeiooa A ~394 I- 319 ,___55, £643 7 :- fi‘ 5' :5. fr '6 f4ao :5. a. 1 9; 11”” ‘e (m 5: ~— 240 ‘2; fl 3' E? q - EL 134 " . 7“2§- ,.,, ‘~—-00 <17 g- ‘3 K “672 ‘1; - 3092 “fl€f:;3060 3039 3020 a [643 gieos .=-l510 —1206 1066 fe:~~ -e——-1005 I996 ‘~=>-901 ‘;~ use ammmoomnw 30311111311“de-ms1q-Z'1-1 ,_, ———-—-———— 1635 Figure 6: 53 Normal Raman Spectra of t-l,2-bis(4- pyridyl)ethylene in CH2C12, the aqueous solution of the dihydrochloride, and the dideuterochloride in D20. EXCITATION WAVELENGTH: 6471 A (Kr+). INTENSIIY . ....Némsézaa.£3.25 9252222.? ... 8o z W 0 3 . n u a 1 ... I e _ r L L .-.IF .1 711.! .11». 1L! _ :1 wng mg: ~23 ~23 ~30 W 2 a“ “e u 22 I. 2 H mm "u H» mm nununflr111¥ mu.u mg" mmw w a. C, > 2 - E --.lllk A. w. m m M. u“ .n n o u .._.~.zm3.3aa: £3.28 33:33.33 5 am: a m. _ I _ m J 7 n m w 11 u n .. m 6 a 1 3 .. a ,7 I. /K\/.lli.i\lu\ .(e. /. ..-Ili rll.(\\, (Qa/ .[IPI'IIL .— b. b . . . F -IFILI mac: “:8 uwcc use .4. 5 A .._.~.zms.3aa._€2.12: ... 62%: H me I _ 1 7 n ‘ e s c _ - .... ,. .... , ... _ 8 2O ._ p ‘:Efilifao >167! {ggzo—"_ / .. ,2 1549 >1492 ‘ 1422 .7141: ‘ 133 1313 1 1 E172 ‘- - -— 993 ;570 954 3300 {cu2cb f»ss7 Isso l I 225 woo 55 mutually exclusive. Since the molecule is made up of two pyridine rings, vibrations in one ring may be in-phase or out-of—phase with the same vibrations in the other ring. In-phase vibrations will be of A symmetry (Ag, Au) i.e. symmetric with respect to the C2 axis. Out-of—phase motions will be antisymmetric with respect to the C2 axis or of B symmetry (Bg, Bu). Analysis of the number of irreducible representations for t-BPE predicts the observation of 23 Ag, 10 Bu' 11 Au' and 22 Bu modes in the vibration spectra. Only gerade modes will remain Raman active. Thus by correlating the symmetry representations of pyri- dine (CZV) modes to the molecular symmetry (CZh) as in Table 2 it becomes evident that each pyridine motion correlates to one Raman active mode and one infrared active mode. Combinations and overtones involving ungerade modes, whose direct products transform as gerade representa- tions, can also be observed in the Raman spectrum of t-BPE. In Table 2 ethylene modes (DZh) are also correlated to the molecular symmetry (CZh)' The vibrational modes associated with the symmetry species in the correlated symmetries 102'103 included in Table 2. Further correla- (CZV’DZh) are tion to possible surface complex symmetries is also examined in Table 2. The choice of axes is consistent with pyridine assignments in reference (104). The crystallographic space group symmetry was not determined in this study. Therefore, correlation to crystal and site symmetries for polycrystal- line Raman spectral assignment was impossible. Efforts 56 .mMDZHBZOU N mqmfifi fl baa .ma .m .ea m mm ne.naa .nm .nm .nom a A//// \\\\V am nea .aa .e .noa .m .nma e”! ea \v1 t t; \. 4 ova .MOH .oma \\‘|‘nr. :4 we .H .ma enma .Mm .4 .1. a a< .naa .na .nem .ma .m mu Emu >mo mcmpmuwm 11momuuweeww111 ImuumEE>m1 1>uumEE>m1 11111111 mmpoz 1111111111 xmamfipwzmommusm amasowaoz cocoamnwmw .momuumeshm meQEOU mommusm manmmmom Op EOHUMHEHHOU HQEUHSM paw ocmamnuomawpmumm vvmmnlmealu mo knuofiewm unasomaofi mnu Op mamawsum can momma mampmuwm map m0 mmmupmaewm mo coaumamuuou "m mqmée 57 0 ~ "moon/m mo oomocu X H a m 5am m9 0 < mm II o mam o9 .m> .-< < \ 4 H) \N) fimp ..a mo LNU zma «mcoawnuo 11mmmuuwaamm111 IwuumEE>m1 1>auoEE>m1 mopoz onmEou 000musm unasooaoz venomounou .QMDZHBZOU N mqmfie 58 to grow adequate single crystals for X-ray examination are underway. Frequency Region 0-600 cm-l Assignments in this region of the Raman spectrum of t-l,2-bis(4-pyridyl)ethylene are particularly difficult. Other than the lattice modes, most of the bands in this region are of very low intensity and quite broad. Lines at 486, and 400 cm.1 have been assigned to pyridine motions 16b and 16a respectively following Spinner.104 The fre- quency at 225 cm-1 has been assigned to a skeletal vibra- tion in accordance with a similar motion in trans- 105 The X-ray crystal structure determination stilbene. is required in order for assignment of modes in the lattice region of the polycrystalline Raman spectra to be made possible. Vibrations 6a and 6b The in—phase (Ag) components of orthogonal pyridine vibrations 6a and 6b appear at 641 cm.1 and 671 cm-1 respectively in the Raman spectrum of polycrystalline t-BPE. Assignment of these modes is fortified by comparison with the same motions in the dihydrochloride. In the Raman spectrum of polycrystalline t-l,2-bis(4-pyridyl)ethylene dihydrochloride the vibration 6b is observed shifted down to 651 cm-1 while vibration 6a appears slightly shifted up to 643 cm-1. The downward shift of 6b frequency is exactly analogous to that observed in the Raman spectra 59 of 4-methyl pyridine and 4-methy1 pyridine hydrochloride as reported by Spinner.102 There, the frequency of the 6b motion in 4-methyl pyridine appears at 669 cm.1 and 651 cm-1 in the hydrochloride. In t-BPE, the behavior of vibration 6a upon formation of the hydrochloride follows that observed for pyridine and pyridinium chloride. In pyridine the Ga vibration appears at 605 cm-1; it falls at 607 cm.1 in 103 a similar shift to higher frequency pyridinium chloride, upon salt formation. A further consideration in the assignment of vibrations 6a and 6b is the relatively weak intensity of 6a relative to 6b in the Raman spectrum of t-BPE and its hydrochloride. In some substituted pyridines and pyridine hydrochlorides the Ga vibration is too weak to be observed102 or appears with very weak intensity. The intensity of 6b is also greater than that of 6a in the Raman spectrum of trans-stilbene (the analogous molecule 105 Regis to t-BPE with benzene rings instead of pyridine). and Corset106 report a 6b/6a intensity ratio of 25/4 in the Raman spectrum of methyl viologen (N,N'dimethyl-4,4' bipyridine). In methylene chloride solution the Ga vibra- tional mode, appearing at 650 cm-1, exhibits a depolariza- tion ratio of 0.17 indicative of a polarized mode as expected for a mode of Ag symmetry. Vibrational mode 6b: however, appears depolarized with a depolarization ratio of 0.64 though the symmetry of this mode is also Ag. The difference in depolarization ratios arises because the vibrational mode 6b transforms as the ayz component 60 of the polarizability tensor (in the coordinate frame designated in Table 2) while 6a transforms asoL , a , YY and azz. Therefore, in the polarizability invariants for XX vibration 6b only the derivative of the yz component of the polarizability with respect to normal coordinate (aayz/an) will be non-zero. The oyz component appears only in the anistropic polarizability invariant and the spherical invariant therefore vanishes, resulting in a predicted depolarization ratio of 3/4. Situations such as this where a totally symmetric vibration is expected to demonstrate a high degree of depolarization, are not common in group theory. In fact the only point groups in which off-diagonal polarizability tensor components transform as the totally symmetric representation are C2h (the case examined here) and the very low symmetry point groups C2, Ci' Cs' and Cl' Vibration 6b cannot effectively couple to the N-H deformation in the dihydrochloride since the motion of the nitrogen atom and hydrogen atom are in the same direction (the form of the normal coordinates for pyridine vibrations are given in Appendix A).107 Thus, no shift is observed upon deuteration of the dihydrochloride with vibration 6b 1 in both salts. Assignments of appearing at 651 cm- vibrations 6a and 6b are particularly important to the understanding of the Raman spectroscopy of t-BPE adsorbed at a silver electrode (see Chapter 6). Comparison of Raman 61 frequencies for modes 6a and 6b in representative molecules is made in Table 2a. Out-of—Plane Vibrations All out-of—plane molecular vibrations appear in the Raman spectrum with a frequency below 1000 cm-1. Only the out-of—phase components of such vibrations are observed in the Raman spectrum since the active symmetry is Bg for out—of—plane modes. A pair of lines appearing in the polycrystalline Raman spectrum of t-BPE at 878 cm.1 and 888 cm"1 have been assigned to pyridine ring vibrations 10a and 10b respectively. The form of the normal coordinate vibrations 10a and 10b as provided by Long and Thomas107 demonstrates that vibration 10a has no component of motion involving the nitrogen atom while 10b involves nitrogen motion (see Appendix A). Therefore, protonation or deuteration of the ring is expected to have little effect on vibrational frequency of vibration 10a. In the Raman spectrum of both t-l,2-bis(4-pyridyl)ethylene dihydro- chloride and dideuterochloride vibration 10a appears unshifted at 879 cm-1. Vibration 10b, in contrast, appears at 894 cm.1 in the polycrystalline Raman spectrum of the dihydrochloride and at 901 cm.1 in the dideuterochloride spectrum. Analogously, pyridine ring vibration 5 has a component of nitrogen motion and can therefore couple to the out—of- plane N-H deformation. In the Raman spectrum 62 TABLE 2a: Comparison of Raman frequencies of modes 6a and 6b in representative molecules. frequencies are in wavenumbers (cm'1)*. Vibrational t-BPE pyridine 4-methy1pyridine t-BPE H2C12 pyridinium chloride 4-methy1pyridine hydrochloride t-BPE D2C12 pyridine deuterochloride 4-methy1pyridine deuterochloride a. Too weak to be observed. b. No available data 641 605 643 607 645 602 671 652 669 651 637 651 651 634 Frequencies reported were those observed with polycrystalline samples for all molecules except pyridine and pyridine deuterochloride. Liquid pyridine Raman frequencies are reported. Pyridine deuterochloride frequencies were observed in D20 solution. all 63 of polycrystalline t-BPE vibration 5 appears with a fre- quency of 958 cm.1 but cannot be found in the Raman spectra of the dihydrochloride and dideuterochloride. The form of normal coordinate 17a (Appendix A) exhibits zero amplitude at the nitrogen position as in vibration 10a. Therefore, 17a is not expected to couple effectively with the N-H out-of—plane deformation and its frequency should be shifted little in the Raman sepctrum of the dihydrochloride and dideuterochloride relative to its position in the t-BPE Raman spectrum. In the Raman spectrum of polycrystalline t-BPE 17a appears at 976 cm.1 and at 970 cm.1 in methylene chloride. In the dihydrochloride and dideuterochloride salts 17a appears at 997 and 996, respectively, demonstra- ting better than a 20 cm.1 shift. However,-in the Raman spectra of the aqueous solutions 17a appears at 977 cm-1 and 970 cm"1 for the dihydrochloride and dideuterochloride: respectively, which is shifted only slightly from the motion in the solution t-BPE Raman spectrum. The out-of- phase components of some of the out—of—plane modes do not appear in the Raman spectrum of t-BPE though they transform as Raman active representations (Bg). Those vibrations absent are v4 and v . Interestingly, these can be observed 11 in the surface Raman spectrum of t-BPE. Ring "Breathing: modes v1 and v12_ The totally symmetric ring "breathing" vibration v1 appears in the vicinity of 1000 cm-1 in all pyridines 64 substituted pyridines,enu1pyridine hydrochlorides. v1 is the most intense Raman line in the spectrum of pyridine and occurs with high intensity in the spectrum of t-BPE. In polycrystalline t-BPE v1 vibration appears at 995 cm-1. In the Raman spectrum of the dihydrochloride v1 appears at 1008 cm-1. The form of the normal coordinate vibration v 107 l nitrogen motion which is directed among the N-H bond in the (Appendix A) shows that the in-plane motion contains hydrochloride. Therefore v1 may couple to the N-H stretch and is expected to occur ataalower frequency upon deutera- tion. In the dideuterochloride of t-BPE v appears at 1 1005 cm—1, which demonstrates the expected shift relative to the dihydrochloride. In the Raman spectrum of pyridine another very intense line appears at 1030 cm-1 and has been assigned to the trigonal ring "breathing" vibration v12. This vibration is conspicuously absent from the Raman spectrum of t-BPE and its chloride salts. The symmetry of this in-plane deformation is Al symmetry in pyridine and thus the in- plane analogue of v12 in t-BPE transforms as the totally symmetric representation and is expected to be observed in the Raman spectrum in the frequency region 1030-1040 cm-l. The appearance of v12 in the surface Raman spectrum of t-BPE and in the normal Raman spectrum of a polycrystalline silver complex with t-BPE is important evidence for the formation of a silver complex at the electrode surface (see Chapter 6). 65 Frequency Region 1050-1300 cm-1 The in-phase component of in-plane ring deformation v18a is very weakly observed in the polycrystalline Raman spectrum of t-BPE at 1069 cm—1. The orthogonal counter- part to vibration 18a, 18b is also expected in the frequency 1 region 1050-1100 cm- but is not observed in the Raman spectrum of t-BPE. 1 At 1184 cm- in the polycrystalline t-BPE Raman spec- trum a very weak line is observed on the tail of the very intense Ceth - ¢ stretch at 1198 cm—1. This weak line is probably a combination or overtone which is in Fermi resonance with the very intense symmetric stretching vibra- tion. If this is the case the symmetry of the combination of overtone must be Ag. All 2nd overtones will be totally symmetric since the result of the direct product of any representation with itself is the totally symmetric repre- sentation. However, the second overtone of a vibration of frequency 592 cm.1 is required and no such vibration is observed in either the infrared100 or Raman spectrum of t-BPE. Interestingly, a very weak line near 600 cm.1 is observed in the surface Raman spectrum of t-BPE which may be the fundamental whose overtone comes into Fermi resonance with the Ce - ¢ stretch. Assignment of this surface mode th is incomplete (see Chapter 6). As mentioned above, the totally symmetric stretch between the ethylenic carbons and the pyridine rings occurs with high intensity in the polycrystalline Raman spectrum 66 of t-BPE at 1198 cm-1. Upon formation of the dihydrochlor- ide and dideuterochloride this vibration shifts to higher frequencies appearing at 1203 cm-1 and 1206 cm-1 in the hydrogenated and deuterated analogues, respectively. The frequency and intensity of the Ceth - ¢ stretch are analogous to that observed in the Raman spectrum of trans- stilbene. Vibrations 9a and 9b appear next in the Raman spectrum of t-BPE. 9a appears at 1225 cm-1 in the polycrystalline spectrum of t-BPE, shifts under the intense neighboring C - ¢ stretch in the spectrum of the dihydrochloride eth and is barely visible at 1215 cm.1 in the Raman spectrum of the dideuterochloride. These observed shifts corres- pond to those observed in the Raman spectrum of 4- methylpyridine. As reported in reference (104), 9a appears in the range 1220-1210 in 4-substituted pyridines and 102 The shift shifts to 1205 cm-1 in 4-methy1pyridine. to higher frequency upon deuteration of the dihydrochloride is also observed in the Raman spectrum of pyridinium chloride. In the hydrogenated salt 9a appears at 1198 cm-1 and at 1202 in the deuterated analogue.108 Vibration 9b is antisymmetric with respect to the C2 axis, transforming as the B2 representation in pyridine, and thus the component of nitrogen motion (Appendix A) is involved in N-H+ in-plane deformation in the dihydrochlor- ide and dideuterochloride. In the Raman spectrum of poly- crystalline t-BPE 9b is assigned to the intense line at 67 1234 cm”1 which shifts to 1254 cm_1 in the dihydrochloride and drops sharply in intensity. Similar to vibration 6b, 9b exhibits a rather high depolarization ratio for a mode of Ag symmetry since the vibration transforms as the yz component of the polarizability tensor as discussed earlier. In the D20 spectrum of the dideuterochloride 9b appears at 1294 cm-1. The shift to higher frequency by 30 cm"1 in the solution (D20) spectrum is an indication of the extensive hydrogen bonding occurring in solution. Frequency Region 1300-1400 cm-1 In this region of the spectrum pyridine ring vibra- tions v3 and v14 appear. The Raman line at 1319 cm-1 in the spectrum of polycrystalline t-BPE is assigned to the in-phase pyridine ring motion v3 in accordance with assignments in other pyridine molecules104 and trans- 105 stilbene. Similar to vibration 9b, the frequency of l v increases upon salt formation, appearing at 1333 cm- 3 in the polycrystalline dihydrochloride. In the Raman spectrum of polycrystalline t-BPE vibra- tion 14 appears at 1351 cm-1 and at 1338 cm.1 in methylene chloride. Vibration l4 correlates to the B2 representation in the pyridine molecular point group as do vibrations 3, 6b and 9b. As for vibrations 3 and 9b, v14 demonstrates a frequency increasing in going to the dihydrochloride and dideuterochloride in the solution Raman spectra. Lines at 1346 cm"1 in the solution spectrum of the dihydrochloride and 1349 cm.1 in the solution spectrum of the 68 dideuterochloride are assigned to v14. On the low frequency side of v in the Raman spectrum of polycrystalline t-BPE 14 appears a vibration at 1342 cm-1. This line is probably the overtone 2x6b which is in Fermi resonance with the totally symmetric vibration 14. The weak line appearing in the Raman spectrum of t-BPE dihydrochloride and dideuterochloride near 1375 cm.1 is probably an overtone of vibration 11 which is of Au symmetry and not observed in the Raman spectrum. In the surface spectrum of t-BPE, however, appears near 685 cm.-1 as discussed in the “11 next chapter. Vibration Pair 19a and 19b Pyridine ring vibrations involving non-hydrogen stretches and bends occur in the frequency region 1400- 1650 cm-1. Vibration 19b is assigned to a weak line 1 in the Raman spectrum of polycrystal- observed at 1419 cm- line t-BPE. Similar to vibrations 6b and 9b, vibration 19b demonstrates the characteristic high depolarization ratio, measured at 0.53. Vibration 3 (B2) demonstrates a very low depolarization ratio (.19) which may be a result of accidental degeneracy with a combination, perhaps 2x6a. As with other vibrations which transform as the B2 symmetry representation in the correlated pyridine symmetry (C2v) 19b couples with N-H+ deformation in the salts of t-BPE. Thus, the vibrational frequency of motion 19b shifts to 1512 cm"1 and 1514 cm-1 in the dihydrochloride and 69 dideuterochloride, respectively. (Further evidence of hydrogen bonding involving the nitrogen proton in solution is evidenced by the shift of the frequency of 19b in the dideuterochloride down to 1463 cm.1 upon dissolution in D20.) On the low frequency shoulder of the 1422 cm-1 line a line appears at 1411 cm.1 in the Raman spectrum of polycrystalline t-BPE. This is another example of Fermi resonance in the spectrum. This mode is probably the totally symmetric combination v4 + v As discussed 11° earlier, neither v11 nor 04 appear in the normal Raman spectrum of t-BPE. v4 is observed weakly at 740 cm.1 in the infrared spectrum of t-BPE.109 v11 is observed at 685 cm'1 in the surface Raman spectrum of t-BPE as mentioned above. Pyridine vibration 19a is observed at 1491 cm-1 in the Raman spectrum of polycrystalline t-BPE. 19a has a component of nitrogen motion along the N—H bond and there- fore is expected to be responsive to formation of the salts. In the polycrystalline dihydrochloride and dideu- terochloride 19a appears as shoulders at 1500 and 1505 cm-1, respectively. Vibration Pair 8a and 8b Still another example of Fermi resonance in the Raman spectrum of polycrystalline t-BPE appears at frequency 1541 cm.1 on the shoulder of fundamental 8b at 1549 cm-1. The overtone is probably 2v11. The out-of-phase component In (M We 70 of vibration 4 is not observed in the Raman spectrum as discussed previously. However, the in-phase (Au) component appears in the infrared spectrum100 at 740 cm-1 and in the surface spectrum of t-BPE (Chapter 6) two lines are observed at 720 cm.1 and 740 cm-1. Thus the Bg out-of- phase component probably occurs at 720 cm.1 and the overtone is expected at 1540 cm-1. Typical of other b vibrations which may couple to N-H deformation vibration 8b, appearing at 1549 cm“1 in the Raman spectrum of polycrystalline t-BPE, shifts upon dissolution of the dideuterochloride salt in D20 from 1606 cm.1 to 1567 cm-1. An analogous shift is observed for vibration 19b. Vibration 8a appears as the most intense line in the Raman spectrum at 1597 cm.1 in poly- 102 observed that 8a shifted crystalline t-BPE. Spinner by nearly 30 cm.1 to higher frequency upon formation of the hydrochloride in the Raman spectrum of 4-methy1pryidine. A similar shift of vibration 8a is observed in the Raman spectrum of t-BPE with formation of the dihydrochloride. Ethylenic C=C Stretch The totally symmetric stretch of the central ethylene carbon atoms appears at 1638 cm-1 in the Raman spectrum of polycrystalline t-BPE. Upon formation of the dihydro- chloride and dideuterochloride salts the carbon-carbon stretch probably overlaps the intense 8a vibration. The weak high frequency shoulder of vibration 8a in the 71 polycrystalline Raman spectra of the salts has been assigned to the C=C symmetric stretch although the low intensity is not completely understood. C-H, N-H(D), Stretching Motions The C-H stretching frequencies appear in the Raman spectrum of t-BPE between 3000 cm-1 and 3100 cm-1. The lowest frequency can be assigned to ethylenic carbon- hydrogen spectrum following assignments in trans-stilbene. Without isotopic substitution studies assignment of other frequencies in the 3000-3100 cm"1 region are ambiguous. In the dihydrochloride aqueous solution spectrum two broad bands centered at 3270 cm-1 and 3400 cm-1 represent N-H symmetric stretching motion. The broadness of these lines is indicative of extensive hydrogen bonding with water. A line appearing at 3266 cm.1 in the polycrystalline t-BPE dihydrochloride Raman spectrum is probably the overtone 2X(C=C sym. str). In the deutrated salt the N-D symmetric stretch appears as a broad doublet with centers at 2400 cm-1 and 2510 cm-1. The polycrystalline dideuterochloride Raman spectrum exhibits no bands in this region. The absence of N-H, and N-D stretches in the polycry- stalline Raman spectra of the salts along with the large shifts in frequency of modes coupling to N-H(D) deformations upon dissolution appears to indicate the relatively weak association of the proton (deuteron) for the nitrogen atom in the crystal lattice. This association is expected 72 to be influenced by the presence of the highly electronega- tive chloride ion nearby in the lattice. The assignments in this chapter are central to an understanding of the Raman activities of t-l,2-bis(4- pyridyl)ethylene adsorbed at a polycrystalline silver electrode (Chapter 6). CHAPTER VI Raman Spectra of t-1,2-bis(4-pyridy1)ethy1ene adsorbed at a polycrystalline silver electrode. Intense Raman scattering from trans-l,2-bis(4- pyridyl)ethylene adsorbed at a polycrystalline silver electrode has been observed in this study. The procedure for electrochemical pretreatment of the silver electrode is described in Chapter 4. Concentrations of t-BPE in initial studies were 2.5 mM in 0.193KC1, but concentrations as low as 250 MM t-BPE produced comparable Raman signals. Normal Raman signals from aqueous solutions of 2.5 mM t-BPE were too weak to be observed. Surface Raman spectra of t-BPE, in contrast, generally demonstrated signal inten- 5 to 3 x 105 counts per second. sity in the range 1 X 10 Under similar electrochemical roughening procedures t-BPE molecules adsorbed to silver from bulk concentrations of 1 mM and above produced a brownish-black coating on the electrode surface while adsorption from bulk concentrations of 500 MM and less produced an electrode coating indis- tinguishable from that observed following electrochemical cycling in aqueous KCl solution alone. Thus studies of surface Raman spectra from t-BPE molecules adsorbed from bulk concentrations in excess of 1 mM are hereafter referred 73 74 to as "heavy coverage" spectra. Those spectra observed from t-BPE molecules adsorbed from bulk concentrations less than 500 MM are referred to as "light coverage" studies. Heavy Coverage Raman Spectra of t-BPE under 5145 A Excitation. The Raman spectra of a heavy coverage of t-l,2-bis(4- pyridyl)ethylene adsorbed at a polycrystalline silver electrode under 5145 A excitation are given in Table 3. Some important sections of the Raman spectra are shown in Figure 6. Compared in Table 3 and Figure 6 are the normal Raman spectrum of t-BPE (assigned in Chapter 5) and the surface Raman spectra of t-BPE recorded at -600 mV versus sce (i.e. relative to the saturated calomel electrode) immediately following the oxidation-reduction cycle, -50 mV versus sce and again at —600 mV versus sce (following recording of the spectrum at -50 mV). Symmetry assignments based on the molecular point group symmetry and possibly surface complex symmetries (C2, C5) are also included in Table 3. In the Raman spectrum of t-BPE on silver recorded immediately after the oxidation reduction cycle (which appears just above the polycrystalline t-BPE Raman spectrum in Figure 7) some interesting differences and similarities to the normal Raman spectrum are observed. The most pronounced difference is the appearance of a new peak at 1 655 cm- in the surface spectrum. Assignment of this Figure 7: 75 Heavy Coverage Potential Dependence of the Surface Raman Spectrum of t-l,2-bis(4- pyridyl)ethylene on Silver. Compared are: the normal Raman spectrum of polycrystalline t-BPE (bottom spectrum); the surface Raman spectrum recorded at -600 mV immediately fol- lowing the oxidation-reduction cycle (above polycrystalline spectrum); surface spectrum at -50 mV; and the surface spectrum recorded on returning to -600 mV (top spectrum). EXCITATION WAVELENGTH: 5145 A (Ar+). cmwll O O N we I I 76 -600 my vs. SCE trans~1.2-bis-(4-pyridyl) ethylene on Ag at «Own.- wmmue. 1 mmefil mme V 1 1 l moelilmluv meal. . mwe MOwhhlllil/JM l. r Smart/L1 - memalil - -J -- coma . mama 23W :2 as m m .1 ll mama m anmmar , aama u m as: mm: ... ... 5 mm: (o 9: gnaw 31-1 mac : 4 r/,-,. aama 3.3.». mm“... Qmmu l. v . t lWLV ms..a...:ll:.v - Eadlr .. - .. Mace—...! .-Hll :-.1 aeea 5 J :55;— -600 my polycrystalline l _l 700 650 L I 1500 1400 1300 1200 1600 (:m'1 77 amm22aezoo m mamE oml um Esnuoomm mcmonooou mGHBOHHOM popuouou Esnuooam + .maoxo GOHUUDUTHIGOHumpme mnu uwumm mawumeoEEH Umpuoowu Esuuoomm .1 k ema nohnnha> mumum UmvHUXQ MOH 3> oem 2> oem 3> oem a ea e am 3> omm 3> omm 3> omm m =4 e we eae mae 3> mme a ea aa aa mee mee mee a .a e:ame he aa mme mme mme a .a he a< mee ne ma ooe ooe oee maa am 3 omm m .a he an 3 amm 3 mm m .< me am 0mm mm mmm omm omm omm mom mom mom omm omm omm oem omm omm omm .nom acuaa mom nhoaa nhoaa nhoaa ama encahnnha> am moahhna me +>s ooe: Il>s cm: >5 ooe: mo no mmmnh a cohaas manoEEmm on am no ma * um m4 mamaaoummuo>aom ucoEGmmmm< Emu co manna co manna no mamnh .wamamhumxameanmaueenah:m.a1h mcaa Iaoumwuowaom num3 mmma GOmmquEou .COHumumoxo < mvam Hopes mponuooam Hm>amm mcmaamummuomaom m ouco UmQHOmpm ocmawcuomaxpmuwmlwvmeIN.H1u mo muuommm cmEmm "m mqmfla 78 Amm22aezoo m mameev .>E omI no Esnuoomm ocapnoomu mCaBOaa0m pmpuoomu Esuuommm + .oao>o coauosponicOaumpaxo may noumm xaouoameEa popuoumn Econommm an cm 3> 3> 3> 3> Nava mmma mNma mama aama ooNa Nooa hmoa mooa Nno vmo mum ovm ...->5 COO-I on em CO mmmlu 3> 3> 3> mava ovma mmma mama onma NVNa ooNa mmoa oooa ovoa mooa mum omm omw >5 Oml on am so mnmuh cm 3> 3> 3 Nava mmma mama NoNa mvNa mNNa mONa maaa vooa mooa oaoa omo mmo owm ovo >E oool ¥ on am SO mmmlu woman wouaoxo e..mu ha . on $2 om .nhn e..o ma hma naa ceaumuoa> oumum UGUHUXG a m mha noa moa Ahhov e> * nowaaz ucmEGmmmm< C‘UWT010 NoumEEN L) m staimcnald E N .QMDZHBZOU m mam<8 79 .>E oml um Esuuoomm ocaouooou ocaBOaaOM Uopnooou Esuuoomm + .oaowo ceauospoHIGOaumpaxo one Hmumm waouMacoEEa popnooou Esnuoomm .4. emea emea emea 4 .4 emea xhhee 4» x m m4 mmma oaoa avoa 4 .< mmma .uum ONO m4 4 .4 aama m o4 aooa mooa moea 4 .< mmma cm 4 emma amma a .4 ha mm mmma mmma mama < .4 mama am 4 4 .4 aama e x m a4 moma oama m .4 oma eaea maea eaea 4 .4 aaea naa a4 emea emea m =4 e x m a4 mvva EOaUMHoa> mumum omuaoxo mmea mmea emea 4 .4 mmea haa a4 +>s ooe: >2 om: e>s ooe: mo no mmmuh # acnaas Nahossxn no m< um o« no o< ocaaamummnomaom ucoficmammd END GO mmmlu CO mmmlu :0 mmmlu .QMDZHBZOU m mamflfi 80 Raman frequency is important to the discussion of molecular orientation at the surface and is reserved until other surface data are presented below. Another distinct difference between the normal Raman spectrum and the surface Raman spectrum of t-BPE, observed in 0.1 M Cl- electrolyte, is the appearance of an intense line at 230 cm-1. Similar lines have appeared in the surface Raman investigations of pyridine5 also and have been assigned to the Ag-Cl symmetric stretch because of its appearance in the surface Raman spectrum even in the absence of pyridine.8 We concur with this assignment by noting that moving the electrode potential negative of the point of zero charge (p.z.c.), which occurs near -0.97 V for silver,102a the intensity of the 230 cm-1 line is attenuated to less than half its intensity at positive potentials and does not recover intensity upon return to positive potentials. The intensities of the other lines in the surface Raman spectrum of t-BPE do not show the same potential dependence. This indicates that, as expected, Cl- ions leave the vicinity of the electrode surface at negative potentials and the vacant sites are filled by rearrangement of the molecules remaining at the surface. Thus, the number of Ag-Cl bonds decreases which accounts for the drop in intensity at 230 cm-1. This is an important point: many researchers in efforts to locate a metal-nitrogen stretch in the surface Raman spectra of l nitrogen-containing molecules have assigned the 230 cm- line to such a motion. 81 Other differences between the normal and surface Raman spectra of t-BPE include the appearance of new weak lines at 685 cm-1, 600 cm-1, 550 cm-1, and a group of three vibrations in the 300-400 cm"1 region. The 685 cm-1 and 600 cm‘1 lines can be assigned to the in-phase (Au) and out-of-phase (Bg) components of vibration 11, respectively, following similar assigments in the vibrational spectra of trans-stilbene.105 The 550 cm-1 line is assigned to the out-of-phase (Bu) component of vibration 6b by analogy to 105 trans-stilbene spectra and its appearance in the infra- red spectrum of t-BPE.100 The low frequency motions have not yet been assigned. A very weak line is also observed at 1118 cm-1 which could be the in-phase (Ag) component of vibration 15 which appears at 1148 cm-1 in pyridine104 but is not observed in the infrared or Raman spectra of t-BPE. Another weak new line observed in the -600 mV surface spectrum occurs at 840 cm-1 which may be assigned to the in-phase (Au) component of model v4 of ethylene which appears at 837 cm-1 in the infrared spectrum of t-BPE.1 Other than the differences mentioned above, the surface Raman spectrum of t-BPE observed at -600 mV under 5145 A excitation after electrochemical cycling is very much like the normal Raman spectrum of t-BPE. Absence of Fermi resonance, intensity drops observed for vibrations v3 and ng at 1317 and 1245 cm-1, respectively, and the shift in 82 the frequency of v6b from 671 cm.1 in polycrystalline t-BPE to 667 cm.1 in the surface spectrum are all analogous to solution effects observedixrthe normal Raman spectrum of t-BPE (see Table 1). Most of the surface Raman lines are only slightly shifted from their counterparts in the polycrystalline spectrum. Upon altering potential to -50 mV versus sce some extraordinary changes in the surface Raman spectrum of t-BPE are observed. The 655 cm-1 line becomes much less intense and the in-phase component of vibration 11 (Au) at 685 cm-1 increases sharply in intensity. The very weak 550 cm-1 line disappears and a very weak line at 539 cm-1 appears which can be assigned to the out-of—phase motion 6a since it appears in the infrared spectrum.109 The intensity of the Raman line at 880 cm.1 in the -600 mV spectrum drops, 1 accompanied by an increase in intensity at 832 cm- in the -50 mV spectrum. The ethylenic v4 motion appears to l O O O C I and increases in intenSIty on mov1ng the shift to 850 cm- potential to -50 mV. The ethylenic carbon to ring carbon symmetric stretch at 1200 cm.1 picks up some asymmetry on its high frequency side accompanied by the appearance of a new line at 1270 cm-1. The in-phase (Ag) ethylenic hydrogen deformation appears at 1329 cm.1 and is assigned by analogy with the corresponding mode in the trans-stilbene vibrational spectra.105 A host of new vibrations appear in the 1400-1560 cm-1 range. Most of these lines can be assigned to the 83 out-of—phase (Bu) ring motions as shown in Table 3. The out-of—phase component of vibrational mode 8b is assigned to the new line at 1554 cm-1 by comparison to that observed in the infrared spectrum of t-BPE where 8b occurs at 1560 cm-1. Similarly, the Raman line at 1510 cm-1 in the -50 mV surface spectrum can be assigned to the Bu component of vibration 19a which appears at 1504 cm.1 in the infrared spectrum. The Raman band at 1480 cm.1 has no counterpart in the infrared spectrum and is probably the totally symmetric overtone of vibration 4 (740 cm-1) which is in Fermi resonance with the totally symmetric component of l in the mode 19a at 1495 cm'l. The line at 1418 cm" -50 mV spectrum can be assigned to the Bu component of mode 19b which appears in the infrared spectrum1 also at 1418 cm-1. A new line appears at 1449 cm.1 in the surface spectrum at this relatively positive potential which has not been assigned and will be discussed later in this section. The totally symmetric ethylenic carbon stretch at 1640 cm”1 is reduced to half the intensity of that line in the -600 mV spectrum. The increased height of the valley between the 1640 cm-1 line and the 1603 cm-1 line with decreased intensity at 1640 cm"1 indicates the possible presence of an isosbestic point accompanying a shift in the frequency of C=C to a frequency near 1600 cm-1. Upon returning to -600 mV the initial spectrum observed at this potential is not recovered. Instead, a spectrum is observed which maintains many of the features observed in 84 the -50 mV spectrum. The in-phase and out—of—phase compon- ents of the pyridine ring motions in the region 1400- 1560 cm”1 increase in intensity and the intensity of the line at 684 cm.1 (V11, Au component) is maintained. The intensity of the line at 655 cm-1 increases, but remains much weaker than that observed in the original spectrum recorded at -600 mV. The in-phase ethylenic hydrogen deformation at 1325 cm.1 increases in relative intensity. 1, 1449 cm"1 and Interestingly, the vibrations at 1270 cm- 832 cm-1 have disappeared from the spectrum upon moving the potential to -600 mV. The intensity of the C=C stretch at 1639 cm.1 has recovered at the more negative potential. It must be noted that the spectroscopic changes observed above occur irreversibly. That is, the original spectrum observed at -600 mV is never recovered once the electrode potential has been shifted relatively positive. Also, and most importantly, these spectroscopic changes are observed only under laser excitation with wavelengths in the green or shorter. Excitation with low energy laser lines (e.g. 6471 A, Kr+) results in the observation of only the spectrum initially observed at -600 mV versus sce and the potential dependence of the surface Raman spectrum is completely reversible. Another requirement for the observation of irreversible potential dependent behavior is that the electrode must have a heavy coverage of t-BPE at the surface. In experiments with bulk t-BPE concentrations at 500 MM and less only reversible potential dependence 85 of the surface Raman spectrum is observed. These experi- mental data are addressed in the following sections. Molecular reorientation and surface complex formation are invoked to explain the anomalous potential-dependent behavior of the surface Raman spectrum. Molecular Reorientation on Complexation of t-BPE at the Electrode Surface. The experimental data presented in Figure 7 and itemized above can be understood by considering the formation of a surface complex between t-1,2-bis(4-pyridyl)ethylene and surface silver atoms. In the surface Raman spectrum of t-BPE observed at -600 mV immediately following the oxidation-reduction cycle the molecule can be considered to be bound to the silver surface through one of the nitrogen lone pairs, i.e. in an "end-on" configuration. This conclu- sion is reasonable in that the attachment of one point limits the amount of electrode density overlap between the metal and adsorbate. Thus, the surface Raman spectrum of t-BPE at -600 mV appears little different from the normal Raman spectrum. The major difference between the normal Raman spectrum and the surface spectrum is the appearance of an intense line at 655 cm-1. The 655 cm-1 line can be assigned to vibration 6b occurring in the pyridine ring closest to the silver surface. When the nitrogen atom is attached to a substituent, vibration 6b always drops in frequency. In t-l,2-bis(4—pyridyl)ethylene 86 dihydrochloride and dideuterochloride 6b is observed at 651 cm.1 (see Table l), in N-methyl pyridium110 at 649 cm-1: and in methyl viologen106 at 657.5 cm-l. The form of 107 vibration 6b according to Long and Thomas requires that the substituent have a component of motion in the same direction as the nitrogen atom and orthogonal to the nitrogen-substituent bond. Thus, some aplitude of silver motion is required in the vibration 6b when t-BPE is attached "end-on" to the surface. This silver motion is therefore available for vibronic coupling of metal electronic states withtflmaincoming radiation and a resonance Raman mechanism becomes a plausible explanation of the ob- served large Raman intensities. In a resonance Raman enhancement mechanism only those modes are enhanced which most effectively couple to the electronic state in resonance. In the "end-on" configuration only one pyridine ring would show electron density overlap with the metal thus enhancing motions in that ring which are effective in vibronic coupling. Thus the 6b motion in the ring closest to the surface (655 cm-1) appears with higher intensity than the same motion in the ring more remote to the surface (667 cm-l). Motions which do not include motion of the nitrogen atom are expected to be absent from the surface spectrum of the "end-on" t-BPE. Thus, vibration 10b (888 cm-1) cannot be found in the surface spectrum of t-BPE at -600 mV and mode 17a appears very weakly at 958 cm-1. A pyridine vibration which is very similar to 87 motion 6b is mode 15. Vibrational mode 15 is not observed in the infrared or normal Raman spectrum of t-BPE but appears very weakly at 1118 cm-1 in the surface Raman spectrum of t-BPE in the "end-on" configuration. Most of the other modes observed in the surface Raman spectrum at -600 mV ) transform as the totally symmetric representation (C2h and may therefore be an indication of a Franck-Condon scattering mechanism (see Chapter 3). On shifting the potential to -50 mV several ungerade modes appear, along with other changes. Ungerade modes appear exclusively in the infrared spectrum of t-BPE thus implicating a reduction in the symmetry of the system probed in the Raman experiment. The "surface selection rule", studied extensively by Hexter and Albrecht91 does not provide an explanation for the appearance of infrared modes in the surface Raman spectrum of a centrosymmetric molecule such as t-BPE (see Figure 2). A reduction in symmetry upon complexation to surface silver atoms, in contrast, adequately accounts for the appearance of infra- red modes. The new line at 1270 cm.1 is assigned to the symmetric ehtylene-carbon-to-ring-carbon stretch in an excited n* state of t-BPE. This conclusion is made on the basis of comparison with the excited state vibrational spectrum of trans-stilbene (observed experimentally by low-temperature, high-resolution crystal absorption spectroscopy of trans- 109 stilbene in a dibenzyl matrix by Dyck and McClure ). In 88 trans-stilbene the Ceth - ¢ symmetric stretch is observed at 1191 cm-1 in the ground state (fluorescence spectrum) and at 1250 cm.1 in the excited state. A line at 1427 cm-1 in the excited state vibrations of trans-stilbene had no counterpart in the fluorescence spectrum and remained unassigned. A similar phenomenon occurs in the surface spectrum of t-BPE at -50 mV. A line at 1449 cm-1 apparently also has no counterpart in the ground state vibrations since it disappears along with other proposed n* Vibrations upon returning to -600 mV. The low-temperature, mixed crystal absorption spectrum of trans-stilbene exhibits a sequency of very strong lines separated by 1599 cm“1 which has been assigned111 to the C=C symmetric stretch in the excited state. A similar shift in the C=C stretch in the t-BPE surface Raman spectrum at -50 mV is invoked here to explain the reduction in intensity at 1640 cm-1. Such a shift would bring the frequency of the ethylene motion under- neath the peak of vibration 8a occurring at 1603 cm-1. The shift in the frequency of mode 10 from 880 cm“1 to 832 cm-1 on moving to relative positive potentials is analogous to a shift observed in going from the fluorescence spectrum to the absorption spectrum of trans—stilbene111 l and 847 cm-1, where lines are observed at 866 cm— respectively. The potential dependence of the lines at 880 cm"1 and 832 cm.1 shown in Figure 8 is clear evidence that these Raman lines are associated with the same Vibration, not different ones. 89 Figure 8: Potential Dependence of Mode 10 in the Surface Raman spectrum of t-BPE "flat" on the silver surface. 4.. EXCITATION WAVELENGTH: 5145 A (Ar ). Potential Dependence of Mode 10b of t-BPE ‘flat' on Ag. . Excitation: 5145A mv SCE 8 00 O in Q '600 N M w M 1 I I 900 800 CM' 91 Thus it appears that vibrations from the n* state in t-BPE are observed in the surface Raman spectrum at -50 mV versus sce. There are two mechanisms which might reasonably account for the appearance of n* vibrations in the surface Raman spectrum. The first is the influence of saturation of the n* state by a high photon density excitation. If saturation occurred, normal Raman scattering from the excited state may be observed. However, these experiments were performed with laser powers of only 50 mW or less. A simple calculation of the number of photons per t-BPE molecule (assuming a minimal surface coverage and a beam waste of 0.5 nm radius) leads to conclusions that, for saturation to be a viable explanation for the appearance of excited state Vibrations, the excited state lifetimes must be of the order of microseconds or larger, which is not likely. The alternative explanation for the appearance of n* vibrations in the surface Raman spectrum of t-BPE is that the n* state becomes significantly stabilized in energy by electron density overlap with the metal. As discussed in Chapter 3, the pentaminebispyridylethylene ruthenium II complex88 demonstrates considerable stabilization of the t-BPE n* state upon complexation such that the n* becomes an occupied state in the complex. In the "end-on" configura- tion at -600 mV there is little overlap of the molecular n system with metallic electron density and the charge on the silver atom is significantly diminished from the charge 92 on the ruthenium II atom. Hence no n* vibrations are observed at -600 mV. Even at -50 mV the charge on the surface silver atoms is still small and under red excita- tion n* vibrations are not observed. Therefore, in order to postulate that the appearance of n* in the surface Raman spectrum is a result of stabilization brought on by overlap of molecular and metal electron densities a reorientation of the molecule at positive potentials must occur. This reorientation must be intimately related to the requirement of 5145 A (or shorter wavelength) radiation since reversible potential-dependent behavior is observed with red excitation. Consider first the t-BPE molecule in the "end-on" configuration at -600 mV. In this configuration electronic eigenstates in the ring closest to the surface can be expected to be perturbed owing to electronic overlap with the metal (particularly those electronic states with non- zero electron density at the bonded nitrogen atom). Those parts of the molecule not experiencing overlap will be little perturbed. Thus of the two n states (i.e. two nitrogen lone pairs) one is expected to be significantly perturbed by interaction with the surface while the other remains fairly constant in energy. For t-BPE in solution the visible absorption spectrum shows a weak n-fl* transition which peaks at 475 nm with a molar extinction coefficient of 1 MI1 cm—1 (Figure 9). This shoulder on the near UV 110 n-n* transition can be assigned to n-n* since in the dihydrochloride the intensity of the visible peak is 93 Figure 9: The Visible Absorption Spectra of I. t-1,2-bis(4-pyridyl)ethylene and II. t-l,2-bis(4-pyridyl)ethylene dihydrochloride. 94 Asjfiueazmmz cos one coo omm com amt ooe _ 1 _ a _ . ‘ am: 5 «2.3.5225: o:o_m._.oa_§:a.3n_h.m.a.a .= .. mam-8 ... 2.23:.a_m_._:e-$n_h.m.a.a ._ Co 930on :o_3..emn< 0.5m; 1 m6 1o.a m.— 1uaiomaoo uonoulixa 1910111 (I-wol-w) 95 attenuated (typical of the behavior observed for n-n* transitions observed in other pyridine compounds upon acidification). The unbonded nitrogen lone pair of t-BPE in the "end-on" configuration at the surface is still available to undergo such an n-n* transition when excited with adequate energy. At 5145 A excitation, as evident in Figure 9, the unbonded ring is slightly in resonance with the n-n* transition. The low symmetry of the surface complex may act to make the transition allowed and increase the actual molar absoptivity. On moving to a relatively positive potential (-50 mV) an electron excited from the unbound nitrogen lone pair into the ethylenic n* orbital would be expected to experience a force of attraction to the positively polarized electrode surface. The coulombic attraction of the n* electron for the positively charged metal surface may cause the molecule to shift its nuclear equilibrium configuration in the excited state. The coordinate along which the shift occurs is that coordinate connecting the n* orbital with the positively charged electrode surface. In general, this coordinate will include contributions from all normal coordinates of the t-BPE molecule with components of motion of the ethylene carbons out of the plane of the molecule. Thus, an electron excited out of the unbound n state to a n* state experiences a force which shifts the equilibrium configuration of the excited state relative to the ground state along the coordinate connecting the 96 n* orbital with the surface. The motion of the unbound part of the molecule out of the molecular plane closer to the electrode surface can be expectedtx>lead to further electron density overlap with the metal. The increased overlap afforded by the molecule "lying down" on the surface can act to stabilize significantly the n* state to within the occupied eigenstates of the silver-t-BPE complex. Thus, the n* state appears as a ground state in the surface Raman spectrum of t-BPE at -50 mV potential under 5145 A excitation. The absence of similar molecular orientation at -600 mV (observed initially following anodization) is accounted for by remembering that at -600 mV the positive charge at the electrode surface is reduced (relative to that at -50 mV). The reduced electrode polarization diminishes the force experienced by an electron in the n* state. Reemission occurs from an excited state only slightly shifted along the coordinate connecting the n* orbital with the metal surface. Thus reorientation is not affected. With t-BPE molecules now "flat" on the surface, and extensive electron density overlap of all sections of the molecule with the metal, more vibrations are expected to couple effectively to electronic transitions between eigenstates of the surface complex electronic manifold. Therefore, more modes appear, with greater intensity, in the surface Raman spectrum of t-BPE by virtue of a resonance Raman enhancement mechanism. Accordingly, the 97 ring motions in the region 1400-1560 cm.1 increase in intensity since the rings are now flat on the surface. Observation of the in-phase ethylenic hydrogen deformation at 1329 cm-1 is further indication that other sections of the molecule experience electron density overlap with the metal at -50 mV. The irreversibility in the potential dependence of the surface Raman spectrum of t-BPE, evidenced by the different spectrum observed on return to -600 mV, is a manifestation of the extensive molecule-metal overlap in the "flat" configuration. The stabilization afforded by the overlap makes the "flat" configuration thermodynamically favored over the "end-on". The disappearance of the n* vibrations on returning to the -600 mV potential implies that only at relatively positive potentials is the n* state sufficiently stabilized to enter the manifold of occupied states. Evidence in the excitation profile of t-BPE at different potentials corroborates the stabiliza- tion of energy states on going to positive potentials (see the following section). Not all t-BPE molecules have changed orientation at the surface, as evidenced by the lack of complete attenua- 1, 1200 cm-1, and 880 cm-1 as well as the 1 tion at 1640 cm- increase in intensity at 655 cm- after returning to —600 mV. Regardless of the number of times the potential is shifted positive (~50 mV), the intensity of the 655 cm"1 line is never completely attenuated. This indicates that 98 the surface can only accomodate a limited number of mole- cules in the "flat" configuration with "end-on" molecules filling the gaps. The enhancement mechanism being proposed in this discussion is one of a resonance Raman enhancement brought about by the increased density of electronic states upon complexation to the silver surface. Evidence that elec- tronic energy distribution in the complex differs from that of the molecule itself can be obtained from the excitation profile of t-BPE on silver. This is discussed in the next section. Excitation Profile of t-BPE on Silver. The variation in the intensity of lines in the surface Raman spectrum of t-BPE with excitation frequency is reported in Figure 10. The excitation frequency profile was recorded in the rhodamine 6G range from 15350 cm-1 to 17350 cm-l. The internal standard to which all intensities were normalized was the 935 cm-1 line of the perchlorate ion (ClO4-) which was added to the supporting electrolyte at a concentration of 0.5 M. The focusing used was point focus and the intensity of the 935 cm-1 could be regulated by positioning the electrode surface at different depths in the electrolyte. The intensities of the t-BPE surface Raman lines are corrected for the fourth power dependence of the scattered frequency (ms) in equation (1) of Chapter 3. Normalization to intensities at one excitation frequency was not carried out since doing so would effectively double Figure 10: 99 Excitation Profile at t-l,2-bis(4- pyridyl)ethylene on silver at -600 mV sce in the end-on configuration (Rhodamine 6G excitation frequency range). Data points are normalized to the 935 cm.-1 line of the perchlorate internal standard and corrected for the 3’4 dependence of the normal Raman cross section. Excitation Profile of t-1,2-bis-(4-pyridyl)-ethylene on Ag at -600 my vs. SCE (end-on). 16000 16500 17000 1 15500 cm" 101 the absolute random error. Therefore the most intense points in the excitation profile arise from the most intense line in the surface Raman spectrum. The excitation profile was examined for a heavy coverage of t-BPE molecules on silver in the "end-on" configuration observed at -600 mV versus sce. Owing to intensity fluctuations with movement of the laser beam to different points on the silver surface precautions were taken to assure excitation of the surface spectrum at the same point on the surface for each excitation frequency. The experimental arrangement is discussed in Chapter 4. The values plotted in Figure 10 and listed in Table 4 were obtained for each Raman line at each excitation frequency by the equation: Iv' -4 -1 18 F = (v ) (1 X 10 ) (12) I935 S where F is the value plotted, IV, is the intensity of the vibrational Raman line at a particular exciation frequency, I is the intensity of the internal standard (C104 ) 935 line at 935 cm-1 and 3s is the scattered frequency in ‘wavenumbers. The absolute random error in the points (F) was calculated from A (AIV.1(US41'1(1 x 1018) (4193511v.(5541(1 x 1018) F: + I935 1335 (458141v.(1 x 1018) + __5 . (13) I935V 102 .aIEo Ga ma mucosoonm coaumuaoxm an .wm comoxm no: 0U muonum Eoocmn muDaOmQ< + aa.eaa mm.am em.ma ae.ma am.ee mm.aa me.ma a.eemma .am am.eoa me.mm am.ea am.ma me.em ea.ma me.ma o.ommma .am em.am me.am aa.aa em.aa mm.aa mm.ma em.aa e.aeama .aa mm.am ma.am ao.ma ma.aa mm.me me.ma ma.aa m.eeoma .ma aa.ee am.mm ma.aa om.aa mm.me em.ma ea.oa m.omaea .ma mm.me aa.am ma.aa me.ma aa.me mm.ma me.ma m.ammea .ea aa.am mm.me aa.ea me.ea ma.am me.ea mm.ma m.ommea .ma me.maa ma.aa mm.aa aa.aa em.mm am.mm em.em e.aeeea .ea me.eea me.maa aa.am am.em me.ma em.am me.am m.aemea .ma ee.mea em.eaa am.em mm.mm aa.aoa am.mm mm.mm m.omeea .ma ma.aoa ea.oa eo.aa mm.am mm.am am.mm oe.em m.ommea .aa ee.mm me.me om.ea me.ea ma.ae aa.ea mm.mm e.aemea .oa mm.ma mm.me em.ma oa.ma aa.ae me.am me.em o.omaea .a om.aa ee.me mm.ma am.ma am.me aa.am mm.mm m.meoea .a mm.ma ae.am ee.ma em.am mm.mm em.mm em.mm m.omama .m em.ma am.mm mm.ma am.am mm.mm aa.mm em.mm a.amama .e mm.ma me.am oa.ma am.mm me.ma aa.am ee.mm m.ammma .m aa.eaa ma.aa me.mm aa.am me.ma aa.am me.ma m.mmema .e ae.mma mm.moa aa.mm me.am mm.aa me.mm aa.ee a.mmmma .m aa.ema me.aoa mm.am ee.om aa.moa ee.mm ma.ae e.amema .m am.mma aa.eaa ao.mm me.am ee.ama aa.ee mm.ae m.ammma .a oeea aoea eema mama aama mooa mme .moeoswonm a GOanmuaoxm 11111111111111111 A Boo mucoswonm amG0aumuoa>11111111111111111 +.A 3 How pmoowuuoov mum msmnw> >E oool he ne>aan no ocmamhhehameanmalevnahum.a1h eo maaeona neahnhaoxm "e mam4m 103 The magnitude of the A quantities in eqn. (13) were AIV, = 11 cm, = i1 cm and A58 = i2 cm—l. Thus, the I935 absolute error in each point in Figure 10 is less than 5%. Tfiuaexcitation profile shows the increased intensity enhancement in going toward red excitation wavelengths as observed in the surface Raman studies of pyridine and other molecules at silver as discussed in Chapter 2. More impor- tantly, however, is that the excitation profile exhibits a peak at 16500 cm.1 which does not correspond to the frequency of any absorption in the visible absorption spectrum of t-BPE (Figure 9). The extinction coefficient of the t-BPE 1 (606 cm-1) is visible absorption spectrum at 16500 cm- near zero. The peak indicates a 0-0 transition from the ground state to an excited state, i.e. from the state of zero vibrational quantum number in the ground state to that in the excited state. This is indicated by the demonstration of a peak at 16500 cm- in the profile of each Vibrational mode intensity. If the transition between Vibrational states of the ground and excited state of vibrational quantum number greater than zero the peak in the profile would be shifted for each vibrational frequency. There is some evidence in the profile of a 0-1 transition also (i.e. a transition from the zeroth vibrational level of the ground state to the first vibronic component of the excited state). This evidence is the appearance of a crossover of intensities of the 1200 cm-1 line and the 104 1608 cm_1 line at 15750 cm-l. This may be the 0-1 com- ponent of the transition which peaks (0-0) in the red not shown in this profile range. If this does indicate a 0-1 1 1 transition then the O-Oieiexpected at 14550 cm- (687 cm- ) which is near maxima observed in other surface Raman exci- tation profiles (Chapter 2). If indeed the crossover at 15750 cm-1 implicates a 0-1 transition, then the 1608 cm-1 line is expected to show some relative intensity increase approximately 400 cm-1 further to the blue. At 16150 cm-1 the 1608 cm.1 profile appears to increase slightly compared to the 1200 cm-1 line which supports assignment of the 0-1 transition. The appearance of a 0-0 transition at 16500 cm.-1 indicates the presence of an excited state not normally accessible to excitation in the molecule itself. The state however may be a complex state generated by electron density overlap between t-BPE and the silver surface. The appearance of an intense 0-0 component and a weak or non-existent 0-1 component indicates that the excited state position along all normal coordinates is little shifted from the equilibrium configuration in the ground state. Generally, to observe Franck-Condon scattering in a resonance Raman spectrum the excited state is required to be shifted along a nuclear coordinate such that upon excitationtx>the excited state a force is experienced. For Franck-Condom scattering a 0-1 transition or transition to higher vibrational quanta is expected to be observed 105 with considerable intensity in the excitation profile. Herzberg-Teller resonance Raman scattering requires no such shift in the excited state to contribute to the scattered intensity. Thus, a 0-0 transition is expected to be the strongest peak in the excitation profile of a Herzberg-Teller scatterer. It can therefore be suggested that the intense 0-0 transition in the excitation profile of t-BPE on the silver surface is a result of Herzberg- Teller coupling of excited states of the surface complex. The appearance of an intense absorption at 16500 cm-1 fortifies the hypothesis presentedixrthis thesis that the distribution of states of the molecule at the surface is significantly altered and increased in density upon forma- tion of a surface complex. The complex formed in the "end-on" configurationcxurbe expected to alter the electronic state distribution of states which demonstrate electron density at the bound nitrogen atom. The peak in the excitation profile observed at -600 mV is observed to be potential-dependent as shown in Figure 11. In this figure the excitation profiles of the 655 cm.1 line in the t-BPE surface Raman spectrum are compared at —400 mV and at -600 mV. The actual intensities (normalized to I935 and 534) are listed in Table 5. The 655 cm-1 mode is the most characteristic line of the "end-on" configuration of t-BPE on silver. It is observed in Figure 11 that the intensity of the t-BPE spectrum at -400 mV relative to the internal standard is reduced compared to that observed at 106 Figure 11: Potential Dependence of the Excitation Profile of the 655 cm.1 (v Line in the Surface 61? Raman Spectrum of t-BPE on Silver. 107 a.:=o. oooha comma ooooa comma a__a.flra/_ \+ /+\7rl_l_ [7.2 I! /_.\:/I, _________a__ /.\+\1_/ ... .....- 1/ 2.. coo./ _ \o Z .35 :3 s at: 1...: ...—5 m3 2: .e 2%... Songs Sages: 32.3.... / (to) 101 paioauoo) AllSN 31m 108 .H E0 cfl wocmsvmum I * .wm mmmuxm no: ow muouum Eowcmu wusHOmn<_ - - -4. '1 -l ‘1ft‘llt ..,.'l I .ll1-\ll Ilrl v1 tin-1|"..“III II'I‘I'I‘I'IIIII’y.II, '..llll.,-|".knl I.I‘:.‘I\l'il . I ilil.l.ellll Ii. I'll-III!!! '.II'I .' I. I‘I.i‘llu.l"lll..' ' [I'll't wN.m m.mvmna >¢.ma H.¢¢Mha .Nm .IIIIIIJ"|yI"||‘;I|v‘III. .Il'l.4.- I ll Ill‘ |||l|||l|ulll mm.OH 0.0mmba mw.mH 0.0mmha .HN ma.aa m.mmHhH mm.aa w.mthH .om mm.ma O.Hmona mm.aa m.wvo>a .ma mm.ma m.omme om.oa N.ommmH .ma em.ma m.ommoH me.ma m.Hmm©H .ma ¢h.mH m.mvhma mm.ma m.omhma .ma mm.ma N.om©©H mm.vm «.mvowa .ma NH.BH N.vmmmH mm.mm m.mvmoa .VH mm.ma m.omvma mm.mm m.omvma .ma wh.mH N.ommmH ow.©N m.ommoa .NH 0H.NN h.hVN©H mm.mm «.mvmwa .HH vm.vm m.HmH®H Nv.vm 0.0meH .oa mm.mm o.mvomH mm.mm m.mvooa .m ©N.ON N.HmmmH «m.mm b.0mmma .m me.ma m.¢mmmd em.mm w.ammma .5 Hv.NN o.mmhma v0.5m m.HmhmH .w mm.mm h.¢mme mo.hv m.mmwma .m mh.mm o.mmmmH mm.vv 0.0mmma .v mm.vm o.vamH mo.wv w.amvma .m mm.mm o.vmmmH mm.oo h.HmmmH .N hm.hN m.mmmma .H :80 mmwv wocmsvmum A :60 mmmv «xocmsvmum le uflmcmucH :oflumufloxm uflmcmucH aoflumufloxm mum m5muo> >8 oovl Amomv >E cowl ' l«|;.|‘"|‘ .I .lll-l-1| ‘-.. "l‘ 1"... II" lll’lllll'gdlllll-A'II'I‘lnllll‘.ll.lt‘lu|.\|‘l£1l‘“ i, . ‘-l"lll 1! l O. +.wom mswum> >8 oovl mam com: um um>HHm co mamahnuwAaxwfiuwmlvvmHQIN.HIu mo Esuuowmm :mEmm mnu CH Anw>v mafia HIEo mmo mnu mo mHHmoum coflumufluxm ucwwcwmma Hafiucmuom um mqmde 109 -600 mV. This is a manifestation of the combined increase in concentration of C104- ion near the more positive surface at -400 mV with the decrease in the number of adsorbed t-BPE molecules on moving to positive potentials. The important point to be taken from the potential dependence of the excitation profile is that the peak shifts to lower energy with the corresponding shift to more positive potentials. The peak in the profile at -400 mV occurs at 16150 cm-1 relative to the 16500 cm-1 peak in the -600 mV profile. The shift to lower energy at positive potentials is consistent with the appearance of the n* state in the -50 mV spectrum of t-BPE on silver (Figure 7) and its disappearance on returning to -600 mV. That is, stabilization is sufficient at positive poten- tials for n* to become an occupied state but not at more negative potentials. Frequently in SERS studies researchers have reported potential dependence of surface Raman intensities and drawn conclusions about surface coverage and other infor- mation essential to analysis of the enhancement mechanism. The potential dependence of the excitation profile of t-BPE at a silver electrode demonstrates that such conclusions can be premature and misleading without knowledge of the shifts in intensity expected by excitation profile arguments. The excitation profiles reported here were determined by continuous variation of excitation frequency with spectra 1 recorded every 100 cm- (~H4Jlnm). In other excitation 110 profiles reported in the literature (see Chapter 2 for references) spectra have been reported in intervals of 500-1500 cm’l. The full width at half height (FWHH) of the peak observed in the excitation profile of t-BPE on silver is approximately 300 cm-1. Therefore, early profiles reported may actually have missed important structure in the excitation profiles and consequently the density of states for adsorbates at a metal surface has not previously been recognized as large. Most researchers have concentrated on explanation of the increased enhancement to the red which can be easily understood if the density of states is high. Light Coverage Studies and the Raman Spectrum of an Ag (t-BPE) Complex. Another requirement for the observation of the irrever- sible potential-dependent behavior observed for the t-BPE surface Raman spectra in Figure 7 is that the adsorbate coverage must be heavy. As mentioned earlier, when the bulk concentration of t-BPE prior to the oxidation-reduction cycle does not exceed 500 ng the adsorbate coverage of the electrode is light and only reversible potential dependence of the SERS spectra is observed. Some insight into the effect of light coverage can be gained by analysis of the potential dependence of the light coverage spectrum of t-BPE as shown in Figure 12. The spectra shown were excited with 6578 A radiation (Rhodamine GG emission) although any 111 Figure 12: Potential Dependence of the Surface Raman Spectrum of a Light Coverage of t-BPE on Silver. EXCITATION WAVELENGTH: 6578 A (Dye Laser Emission). 112 35...... ...—.2538 ... .. E... 833“... a 7r»....m......E.E....=i§. .... R. Q 323......" «mun» _ 1200 2D “‘3 ‘ :‘ f" 1013 3655 ~_-% - .TD 1839 ‘__> D ..3 —“‘ 1&4 .2) .) (cw 93S «3.552 3...: 113 wavelength excitation of light coverages produces only reversible potential dependence of the observed surface spectrum. An interesting feature of the potential depen- dence shown in Fig. 12 is the dynamics of the 655 cm-1 . . . . . -1 line. On mov1ng to p051t1ve potentials the 655 cm line decreases in intensity and shifts to higher frequency, coalescing with the 664 cm.1 line (v ). This strengthens 6b the previous assignment of the 655 cm-1 line to motion 6b in the ring attached to the silver surface as discussed in an earlier section of this chapter. A similar b-type mode, 9b, observed at 1244 cm.1 in the surface spectrum exhibits similar potential-dependent behavior, decreasing sharply in intensity on going to positive potentials. This and other changes in the surface Raman spectrum of a low coverage of t-BPE can best be understood by com- parison withtflmenormal Raman spectrum of a polycrystalline silver complex of t-BPE shown in Fig. 13. The stoichio- metry of the complex has not yet been determined but some conclusions can still be drawn in the comparison. In the polycrystalline silver complex spectrum the 1201 cm.1 line (C h - ¢ symmetric stretch) appears at the same intensity et as the v1 mode at 1025 cm-1. Similarly in the surface spectrum of t-BPE at -100 mV versus sce the two vibrations appear with similar intensity. Two lines appearing at 1037 cm.1 and 1040 cm-1, probably split by site symmetry in the crystalline matrix, in the polycrystalline complex spectrum can be assigned to the pyridine ring vibration v12 114 Figure 13: Normal Raman Spectrum of a Polycrystalline Silver Complex at t-BPE (Agn (t—BPE)m) . EXCITATION WAVELENGTH: 6471 A (Kr+). 115 Polycrystalline Silver Complex of t-1.2-bis(4-pyridyl)etllylene [02! ~'—4~—_._.._l {[91 1 200 1 400 J - 600 1 800 cm“ 4 1200 1000 1400 1600 3100 3050 116 ..mezHBzou v mamme. .coHusHOm NHUNSU cH ommH um muoooo coHumubH>e musmHm .um>HHm co mcmecumAHmpHummlvvanIm.HIu wo uon HMHucmuoqucmuuoo mzu CH 858me8 HmooH mo ucHom + .NH. mnomHm cH popuoowu mum muuommm womum>oo ubmHH one .4. HmMH mmmH ommH ¢H mHMH mHMH mmmH m «mmH mmmH mva bm *mNNH mmmH vmmH Mm mmHH anH HomH .uum e I U mmOH moOH vBOH omH ovOH NH VMOH NMOH mNOH H mom MHoH mHOH Ham eem no. mum mmm mmm mOH Hue moo vow mmm no mmm .hum ZIm< ovm omm vmm .HQH> Hmummem omm .uum Huumm HNH MQOHH mvH mm mopoz Hm moHuumq me mm mmqu .wom. >8 ooea um E.mmmuu.cm¢ #:OmHHz mcHHHmummuowHom um>HHm co mmmlu mcHHHmummuo>Hom uswEQmHmm¢ Illf .I..|Il \...I. ‘1’ I'll. I, ll Cl II .ESHuoQO cm8mm mmmn Iu mcHHHmummuoxHom on» on cam .mom >8 oovn um .ewmmuw>oo uanHv Hw>HHm co mcmHhcuwfiprHummI avaQI N HI u mo Eouwommm cm8mm may ou mCOmHHMQEOo pom wcmH>£um IAH>©HH>QI vvaQI m HI u mo waQEOo um>HHm mcHHHmuwwnowHom m mo Eouuommm CMEmm no mqmée 117 «mom mmom «mom OMHm coumupm MHom owom mIU mmoH vmmH oqu .upm UHU hmmH mooH MHoH om mmmH v + on mva mmmH mmmH am HmvH quH mmvH m.mH NNvH HmvH vmvH an mmmlu .mom. >5 ooen um E.mmmuu.cm< e comHHz wCHHHmHmmuowHom Hm>HHm so mmmlu mcHHHMHmwuomHom #:08cmHmmfi . DMDZHBZOU o mqmflfi 118 which occurs in the Raman spectrum of pyridine at 1037 cm-1 and is the most intense line in the spectrum. Vibration 12 is conspicuously absent from the heavy coverage spectrum but appears weakly at 1035 cm.1 in the light coverage spectrum, reaching maximum intensity at the -400 mV potential. The appearance of v12 in the light coverage surface Raman spectrum of t-BPE and its appearance in the Raman spectrum of the polycrystalline silver complex strengthens the hypothesis of the existence of a surface complex. The C=C symmetric stretch at 1639 cm—1 attenuates in intensity going to positive potentials and us currently not completely understood. It seems that most of the scattering observed in the light coverage originates in the part of the molecule closest to the surface. Thus, the Ce - ¢ symmetric stretch (1200 cm-1) and the C=C th symmetric stretch are attenuated at positive potentials and the v1 vibrations (1013 cm-1) appears to narrow. The reason for more intensity arising from the closer part of the molecule can only be decided when the state in which the excitation wavelength is in resonance (assuming a resonance Raman enhancement mechanism) becomes known. There are several electronic excited states which, by using typical building up principles, exhibit electron density, only at the rings and not in the ethylene bridge. Resonance Raman excitation to such a state would be expected to result in increased intensity for the bound ring only. The attenuations of the above mentioned line intensities may 119 not be observed in the heavy coverage because of intense scattering from multilayers. Although some of the features of the light coverage surface Raman spectra are not completely understood other features can be utilized to understand the requirement of heavy coverage for the observation of the "flat" configura- tion. At positive potentials the silver surface atoms have developed nearly a unit positive charge per silver atom. The silver atom in the polycrystalline silver complex is, of course, formally in the +1 oxidation state, experien- cing a full positive charge. Withealight coverage of t-BPE molecules at the surface, chloride and perchlorate ions are expected to occupy more surface sites. The increased concentration of anions at the positively polarized surface acts to screen the positive charge seen by the neutral t-BPE molecule at the surface. The screening effect is invoked to account for differences between the light coverage spectrum and that observed for heavy coverages and for the polycrystalline silver complex. Modes with components of motion orthogonal to the direction of the attraction force at the surface (i.e. those modes in the end-on configuration which are orthogonal to the surface normal) and are not expected to be influenced by charge fluctuations at the surface. However, modes with components of motion along the direction of the attractive force are expected to be influenced by charge fluctuation 120 at the surface. Modes which involve asymmetric stretch along the surface normal (end-on configuration) demonstrate the largest sensitivity to surface coverage. Examples of this are now discussed. Vibration 9a, which appears at 1234 cm.1 in the polycrystalline silver complex Raman spectrum demonstrates asymmetric stretching along the surface normal. This vibration appears at 1233 cm.1 in the heavy coverage surface Raman spectrum of t-BPE at -100 mV but at 1225 cm-1 in the light coverage spectrum. Similarly, vibration 19a appears at 1498 cm.1 in the polycrystalline silver complex spectrum and at 1497 cm-1 in the heavy coverage spectrum of t—BPE on silver at -100 mV. The motion 19a appears at 1493 cm-1 in the light coverage spectrum shown in Figure 12. These observed shifts and the shifts of other a-type modes in the light coverage spectrum relative to the heavy coverage and complex spectra are therefore understood by considering the reduced positive charge seen by the light coverage of t-BPE molecules owing to anion screening. This screening is undoubtedly responsible for the lack of reorientation of the t-BPE molecule in a light coverage under 5145 A irradiation. As discussed in the previous section, reorientation of the t-BPE molecule occurs through excitation into a n* orbital which, owing to the proximity of the positively charged surface, experiences a force along the coordinate connecting the n* orbital with the silver surface. With the positive charge on the electrode screened 121 by the presence of large numbers of anions in the light coverage situation the force exerted on the w* electron is reduced. The reduction in the force is apparently sufficiently large enough such that reorientation cannot be affected. Thus, under light coverage conditions, the "flat" configuration never materializes (at any excitation wavelength) and completely reversible potential-dependent behavior of the surface spectrum is observed. The effect of exciton splitting of an excited state of one t-BPE molecule relative to the excited state of its neighbor on the surface may also account for the observed coverage-dependent phenomena. However, physical evidence for both the exciton model and the screening model are currently lacking and further study is needed. Negative Potential Results and the Reduction Product of t-BPE . Interesting potential-dependent spectroscopic phenomena are also observed on moving to potentials negative of -600 mV. As shown in Figure 14, the surface Raman spectra of t—BPE develops features characteristic of the "flat" configuration. This is observed by comparing the "end— on" spectrum at -400 mV with spectra observed on moving to -1.0 V versus sce. In the spectrum observed at -800 mV intensity has developed at the 681 cm-1 line (vll' Au) 1 with decreasing intensity at 655 cm— . This indicates that the "end-on" configuration is diminishing in Figure 14: 122 Potential Dependence of the Surface Raman Spectra of t-BPE Potentials -400 mV, -600 mV, -800 mV and -1.0 V versus sce. EXCITATION WAVELENGTH: 5145 A (Ar+). 123 - 1639 “C=C on I; 0.5! cm;- . >16“ at -1.0 V v5.50! ' > 1342 1320 IIIEISIIV 1 “ 600 .v J VMW‘ wry), - 400 IIV ‘i J Wu” cm“ £00 1 | ‘ 5- 1010 ...“,H __._ l____J - ,.____1_____ 1600 1500 1400 1300 1200 1100_ "1000 124 prominence and the "flat" configuration takes its place. The increase in intensity of the vibrational lines in the region 1400-1560 cm.1 is further indication that the rings are beginning to lie down on the surface and experience electron density overlap with the metal. The increased 1 intensity at 1320 cm- (C - H in-place, in-phase, eth deformation) is consistent with that observed in going to the flat configuration. By comparing the relative intensities of the C=C symmetric stretching vibration at 1639 cm.1 and the intensity at 1200 cm-1, it is apparent that the 1639 cm.1 band becomes attenuated relative to the 1 C - ¢ stretch at 1200 cm- . This may be indicated of eth the electronic state from which the Raman signal is origi- nating from. At potentials near the p.z.c. (point of zero charge) of silver (i.e. near -900 mV sce in 0.1 g Cl-) it might be expected that electronic excited states would become occupied by transfer of electron density from the metal to the adsorbate. The lowest such excited state for t-BPE can be determined by using known electron density distribu- 111 and tions for the two rings and the ethylene bridge applying building-up principles. The most probable lowest excited state is one of B9 symmetry which therefore can contain a contribution from the ethylenic n* orbital which transforms as Bg symmetry in the molecular C2h point group. In this state there is no electron density at the ring carbon involved in bridging to the ethylene carbon. Thus, 125 no overlap will exist between the W* orbital and the ring carbon at the 4-position. The force constant of the Ceth - ¢ bond is therefore not expected to be different from that in the ground state. Accordingly, the Ceth - ¢ symmetric stretch at 1200 cm.1 is not attenuated by a shift to another frequency even though the system is in a n* state. The C=C symmetric stretch, however, is expected to be attenuated since the force constant of the bond between the ethylene carbons will change and move under the sym- metric ring stretch near 1600 cm_1. A new peak appears in the negative potential surface Raman spectra of t-BPE at 1291 cm-1. This mode can be assigned to the ethylenic hydrogen out-of-phase, in-plane deformation (Bu) because of its appearnace at 1289 cm-1 in the infrared spectrum of t-BPE as observed by Katsumoto.l Interestingly this mode does not appear in the spectrum of the "flat" configuration at -600 mV as shown in Figure 7. This may indicate that the state bound in the "flat" configuration is a state which contains only in-phase electron structure (i.e. is of A type symmetry) while the state observed on sweeping to negative potentials is probably an out-of-phase state as denoted by the B9 symmetry assigned above. The reasons for the molecular reorientation at negative potentials are similar to those for reorientation under 5145 A irradiation at -50 mV (Fig. 7). With the occupation of a n* orbital at negative potentials reorientation may 126 occur in order that better adsorbate-metal overlap be affected. The force experienced by the w* electron in the vicinity of the still relatively weakly positive electrode surface is undoubtedly small and accounts for the minimal amount of reorientation observed spectroscopically. Reori- entation of the end-on molecules is further facilitated by decrease in surface concentration of anions from the supporting electrolyte. The surface sites vacated by anions at negative potentials can be filled by molecular reorientation of the neutral t-BPE molecule at the surface. At potentials negative of the p.z.c. (i.e. -l.0 V versus sce) anomalous spectroscopic changes are observed which further indicate electron transfer from the metal to the adsorbate. The totally symmetric ring motion (8a) at 1604 cm.1 becomes attenuated in intensity accompanied by appearance of a new line at 1560 cm-1. The high fre- quency side of the C=C symmetric stretch at 1639 cm.-1 develops a shoulder in the surface Raman spectrum of t-BPE at -l.0 V. Some distortion in the 1240 cm-1 range is also observed. These changes can be best understood by continuing to alter potential in the negative direction. At -1.2 V versus sce an obviously distinct surface Raman spectrum of t-BPB is observed (Figure 15). In Figure 15, the 1561 cm.1 peak, which was developing in the spectrum at -l.0 V, has become the most intense line in the spectrum. This line is assigned to the totally symmetric pyridine ring motion 8a in the "new" molecule by Figure 15: 127 Surface Raman Spectrum of the Two-Electron Reduction Product of t-BPE on Silver at -l.2 V versus sce. EXCITATION WAVELENGTH: 5145 A (Ar+). 128 .....e GOG GOG." O0: OONH OOMH 09v.— OOmH cow.— OONH, _ _ _ _ _ _ _ H _ 299! ....3 3.2.3.... €2.88 ... .... .2. .. w. .. . . .... . l e ...Olwl. _. I99! .. 3H L! r n — l; AIISNJINI 129 TABLE 7; Raman Spectrum of the two-electron reduction product of t-l,2-bis(4-pyridy1)ethylene on silver at -l.2 v versus sce. Excitation: 5145A. Vibrational Frequency Suggested Assignment (Wilson #) (cm’ ) 641 (6b) in—plane ring deformation 871 10a out-of-plane CH deformation 1005 (1) symmetric ring stretch 1071 (18a) in-plane C-H deformation 1085 (18b) in-plane C-H deformation 1150 (15) in-plane C-H deformation 1196 (9a) in-plane C-H deformation 1214 ? 1232 ? central carbon symmetric stretch 1245 (9b) in—plane N-H deformation 1287 out-of-phase Ce-H deformation 1330 14 ring stretch 1403 19b ring motion 1440 ? 1507 (19a) ring stretch 1540 (8b) ring stretch 1561 (8a) symmetric ring stretch 1593 ? unreduced t-BPE vibration 1630 ? unreduced t-BPE vibration 1652 Ce-C¢ stretch “_ 130 reason of the attentuationcxfthe 1604 cm-1 line at -1.0 V accompanying the growth at 1560 cm-1. It is clear that the very intense Ceth - ¢ symmetric stretch occurring at 1200b_mzuh.2_ cm" 136 TABLE 8: Raman Spectrum of Monoprotonated t-l,2—bis(4— pyridyl)ethylene on silver at -50 mV and -600 mV at pH 5.6 and at -600 mV after base (OH’) is added. Excitation: 5145 A. Region: 1000-1700 cm-l. Monoprotonated _ Assignment t-BPE on Ag Base (OH ) Added Wilson # —50 -600 mV -600 mV 1 ‘ 1017 1006 1012 18a 1063 1056 18b 1100 1095 Ce-¢ sym. str. 1200 1199 1201 9a 1222 9b (NH def) 1242 1243 1246 Ce-—H def.(out-of 1287 1286 3 phase) 1316 1322 14 1340 1338 1338 4 + 11 1414 1409 19b 1427 1424 1422 19a 1495 1490 1491 8b 1545 1538 1543 8b (outsof- phase) 1555 8b (protonated ring) 1582 1574 8a 1607 1603 1606 Ce=Ce sym.str. 1639 1637 1639 _.* -__ .. -——— _—.._v_ H. 137 1605 cm-1 in the dihydrochloride. The increased intensity of mode 18b at 1100 cm"1 in the surface Raman spectrum of t-BPE at pH 5.6 can be understood in that 18b is expected to couple to N-H deformation. Thus it appears that the t-BPE molecule on the electrode surface is monoprotonated at pH 5.6. This conclusion is not unreasonable. In the end-on configuration the t-BPE molecule has one nitrogen lone pair extending into the electrolyte solution. Lavalee and Fleischer88 report values of 5.92 and 4.49 for the pK2 and pK values (negative logarithm of the acid dissociation l constants), respectively, of t-BPE. Though the acid dissociation of the complex (t-BPE at the silver surface) is expected to be somewhat different from the dissociation constants in the ligand, the difference is probably small owing to several factors. In the pentaminebispyridylethy- lene ruthenium II complex the pKa is observed to be 5.0 i 0.1.88 For t-BPE on silver the charge built up on the silver atom is significantly lower than that on the ruthenium atom in the +2 oxidation state. Also, as discussed earlier the increased concentrations of electrolyte anions at the surface are expected to screen the full positive charge on the electrode surface. Thus it is expected that the pKa of the t-BPE molecule—silver surface complex will not be significantly lower than the pK2 of the t-BPE ligand alone. Future surface titration experiments should clarify this point. 138 Some further evidence that the surface spectrum observed at pH 5.6 is that of the monoprotonated t-BPE molecule is presented in Figure 17. There the potential dependence of the pH 5.6 surface spectrum and the spectrum observed after addition of base (KOH) are shown. The appearance of the characteristic lines of the protonated ligand (i.e. 1574 cm-1, 1095 cm.1 and other changes) at -600 mV as well as at -50 mV shows the insensitivity of the pKa value on the true polarization at the electrode surface. The disappearance of the lines, characteristic of the monoprotonated t-BPE-silver complex, upon addition of base is very convincing evidence for the protonation phenomenon at the silver surface. Returning to Figure 16 briefly, some interesting exci- tation frequency dependence for the monoprotonated spectrum is discovered. The lines characteristic of the monopro- tonated species are absent under red excitation but grow in intensity with increasing frequency of the exciting line. This is consistent with the notion that for light surface coverage at positive potentials most of the light scattering intensity originates in the parts of the t-BPE molecule which are closest to the silver surface. The intensity from the ring more remote to the silver surface increases only as a consequence of coming into resonance with a weak n+w* transition in the visible absorption spectrum of both t-BPE and its dihydrochloride (see Figure 9). This evidence also corroborates the hypothesis that the n+n* transition from 139 Figure 17: Surface Raman Spectrum of Monoprotonated t-BPE on Silver at -50 mV (sce), -600 mV (sce), and at -600 mV (sce) after Addition of Base (of) . EXCITATION WAVELENGTH: 5145 A (Ar+). 140 lNTEllSITY Monoprotonated t-BPE on Silver. Excitation: 51451 1201 i E I s i 1700 1500 1500 1400 1300 1200 1100 1000 141 the outside ring can still occur though the distribution of electronic states in the ring closer to the silver surface is significantly perturbed from the distribution of states in the molecule in solution. That the electronic states of the section of the t-BPE molecule not experiencing electron density overlap with the silver surface in the end-on configuration, are n93 significantly perturbed is a requirement for the reorientation mechanism proposed in earlier sections of this chapter. The lines which have been interpreted as characteristic of the monoprotonated t-BPE molecule on the surface also appear in the heavy coverage spectrum observed at -50 mV under 5145 A excita- tion i.e. the spectrum of the flat molecule containing n* vibrations. This spectrum is shown in Figure 18. An interesting experimental fact observed in the studies of the monoprotonated surface spectrum of t-BPE on silver is that upon addition of sufficient acid all surface Raman signals, identifiable as originating from t-BPE molecules, disappear. It is not difficult to understand this fact. A doubly—protonated t-BPE molecule is quite positively charged and is not likely to approach the positively polarized electrode. Also protonating both nitrogens eliminates the possibility of end-on complexation to the metal surface. Influence of C10,- on the Surface Spectrum of t-BPE. The addition of 0.5 N perchlorate for purposes of use as an internal standard in excitation profile experiments 142 Figure 18: Surface Raman Spectrum of Monoprotonated t-BPE on Silver at -50 mV (see) in the "flat" 1 Configuration. 4.. EXCITATION WAVELENGTH: 5145 A (Ar ) 1143 t-BPE on Silver at - 50 mv SCE, pH5.6, 1” 51451 . N o o l 600 l 700 l 800 058 I 88 l 1000 810! z 0901 '90! 1.60! — 1100 l l l 1400 1300 1200 1 1500 (.09! 1600 ‘ Iusuaml 1700 cm 144 has some important influences in the Raman spectrum observed for t-BPE on silver. In general, the screening of the positive charge developed at the silver electrode surface is invoked as responsible for the effects produced by the presence of ClO4-. For the potential dependence of the t-BPE surface spectrum in the negative potential range effects observed in simple 0.1 g chloride electrolytes are 4 to negative potentials t-BPE molecules in the end-on posi- exaggerated with 0.5 5 C10 present. That is, on going tion "lie down" at more positive potentials with half molar perchlorate ions present than in their absence. In both 0.5 I_I 0104’/0.l I_I Cl" and 0.1 I}; 01" alone electrolyte solutions when altering potential in the positive direction the surface Raman signal is attenuated owing to the dis- placement of adsorbed t-BPE molecules by anions at the metal-solution interface. This reduction in surface Raman intensity at positive potentials is understandably augmented by the presence of 0.5 g C104-. In the spectrum of t-BPE "flat" on the silver surface some very interesting influences of the 0.5 g C104- presence are observed. After initiating the molecule reorientation of t-BPE molecules from the end-on configuration to the "flat" configuration at -50 mV under 5145 A excitation the potential was moved to -600 mV and the spectra shown in Figure 19 were observed. In 0.1 g Cl- only the spectrum at ~600 mV typifies what has been assigned to the "flat" spectrum in earlier sections of this chapter. Relative 145 Figure 19: Influence of 0.5 g ClO4 on the Observed Surface Raman Spectrum of t-BPE "flat" on the Silver Surface at -600 mV (sce). EXCITATION WAVELENGTH: 5145 A (Ar+). 146 INTENSITY "“F:>1039 1001 1554 1530 CH- 1480 __;>1494 1425 Influence of cro; on the spectrum of t-l,2 -bls-(4-pyrldyll_-ethylene 'flat' on A: at -600 Mi! vs. SCE. Excitation: 5:45 A 1325 1330 230 1242 .SM era; .1» cr § .1 M cr only l l l“. “a ‘ ' A .N J I l V 2 .I' f 2 I / v a i v w l ' l l 'J l 0\ fl l' ! I! [I \‘ l ‘l V V g ‘ I" I //,// V/ \\.H// 1 wfl\ \J l l l I l A l l 1 n l \J l 1500 1500 1400 1300 1200 1100 1000 147 to that spectrum, in the spectrum observed at the same 4 differences are evident. The intensity of the ethylenic potential with 0.5 M C10 also present some distinct C=C symmetric stretch at 1639 cm"1 remains attenuated in the -600 mV spectrum in 0.5 M ClO4-/0.1 M C1- while regain- ing intensity in 0.1 M Cl- alone. Also, noticed in the perchlorate spectrum is the appearance of a line at 1288 cm.1 which also appears when the t-BPE molecules lie down at negative potentials (earlier section) and which has been assigned to the out-of-phase component of an in-phase ethylenic hydrogen deformation. The weak line at 1120 cm-1 is probably the pyridine ring motion 015. The reason for these differences could rest in the idea that the state that is observed is the highest occupied molecular orbital (HOMO). The HOMO in the case with 0.5 M C10 - present may be different from that when only 4 0.1 M Cl- present as a result of screening. The high concentration of anions diminishes the effective charge at the surface and thus stabilization of different states may result. The state observed in the 0.5 M ClO4—/0.1 M Cl- solution is probably the same state that begins to appear at negative potentials with the t-BPE molecules in the end-on configuration. That is, a H* state which involves no overlap of electron density between the n* orbital and the closest pyridine carbon atom which consequently causes no shift in the Ce - ¢ stretch (1200 cm-1) but shifts th the C=C stretch to lower frequency, attenuating the line at 148 1639 cm-1. The spectrum observed at -600 mV with only 0.1 M C1- present may involve electron density mixing with an unexcited n state since neither of the lines at l 1639 cm- and 1200 cm.1 are shifted or attenuated in intensity. Photodegradation of t-BPE Multilayers on Silver. In the observation of the excitation profile of t-BPE on silver some anomalous frequency dependence is observed. Figure 20 demonstrates this anomalous behavior. The intensity of pyridine ring vibration v11 appears to increase in intensity with increasing excitation wavelength. Ini- tially this phenomenon was attributed to the distinction of the red state appearing in the excitation profile. That is, it was thought that v11 was effective in vibronic coupling to the red excited state but not to other higher energy excited states. However, further investigation proved these assumptions erroneous. The 683 cm"1 line (v11) is absent from the surface Raman spectrum of a light coverage of t-BPE molecules at the electrode- electrolyte interface. The spectra recorded in Figure 20 were obtained with a heavy coverage of t-BPE molecues on the surface. Also, as shown in Figure 21 the intensity of the v vibration is photosensitive to red excitation. In 11 Figure 21 a series of consecutive spectra were recorded in the 625-700 cm_1 region. The only variable changing in this experiment was the time the sample was irradiated. It 149 Figure 20: Excitation Wavelength Dependence of the v11 (683 cm-1) Vibration t-BPE on Silver at -400 mV versus sce. INTENSITY Excitation Wavelength Dependence of trans-l,2-bis-(4-pyridyl)-etlrylene on Ag at -400 m vs. SCE 936 (cm) >>%%%% I 6431 jk sssojk cm" 151 Figure 21: Photodegradation of Multilayers of t-l,2-bis(4-pyridyl)ethy1ene on Silver: Loss of Intensity at 684 cm-1. 152 4 2.... 5.5.5... «523...... mum mflw mflm mum mum . _ . H . <9? i é 9 9 9’ S S 9 8 .7 ..Ill... .....o em a «Re 2.22.4.3.“ ....” ...... .... o3. ... we .... o..o~....o...........-$1...- «.... ... 2o 23...... ... 5:233:32... AIISNJINI 153 1 is clear from the figure that the 684 cm- line is attenu- ated in intensity with increasing irradiation tinmh However, the neighboring vibrations 6b at 667 cm.1 and 655 cm”1 retain full intensity throughout the time of the experiment. The phenomenon pictured in Figure 21 occurred over a time range of 30 minutes at an incident laser power of 50 mW. Obviously, there is evidence of photodegradation but the signal from the molecules at the surface is not degraded. The degradation must occur in the multilayers of t-BPE molecules at the surface. This can be deduced from the absence of the 684 cm-1 line in the light coverage spectrum and the absence of degradation of the characteristic surface mode at 655 cm-1. The photodegradation must necessarily be accompanied by a visible absorption. Since the 684 cm.-1 line increases in intensity in the red so too must the absorption be in the red. But the t-BPE molecule has no visible absorption spectrum above approximately 600 nm. However, in the multilayers the rapid adsorption of t-BPE molecules to the electrode surface probably produced a very low site symmetry seen by each t-BPE molecule. The site symmetry in the multilayers is expected to be C1. Therefore, just as the symmetry of the one-sided attachment of t-BPE molecules to the silver surface is liable to make forbidden transitions allowed so also will the low symmetry in the multilayers. The actual transition could also be an intermolecular one which for stacked flat molecules is likely to occur along the stacking direction. 154 An out-of-plane mode, such as pyridine vibration v11: has nuclear coordinate components only in the stacking direction and may therefore be expected to couple effectively to intermolecular transitions. The visible signcflfphotodegradation in the multilayers is the appearance of a burn hole in the surface coating where the laser beam has been focused. The absence of attenuation in the Raman signals from the molecules directly associated with the silver surface (denoted by the characteristic 655 cm-1 line) is probably a consequence of the ability of the metal substrate to serve as a heat sink, offering an alternative excited state decay pathway to decomposition. The appearance of surface Raman signals from the multilayers should be treated as a caution signal in interpretations. Rowe et a158 and other researchers have reported surface enhanced Raman signals from multilayers. The possible influence of site symmetry lowering of the molecular symmetry on the absorption of incident radiation was not addressed. The observation of surface enhanced Raman signals from absorbed molecules not in direct contact with the metal surface is inconsistent with the proposed enhancement mechanism of surface complex resonance Raman scattering offered in this thesis. CHAPTER VII Surface Enhanced Raman Spectra of Amino Acids and Future Studies. Surface enhanced Raman scattering has been shown to be a versatile probe of dynamics at a metal-dielectric inter- face. For t-l,2-bis(4-pyridyl)ethy1ene adsorbed at a polycrystalline silver electrode nearly every vibrational mode of the molecule, both infrared and Raman active modes, has been observed in the SERS spectrum. Molecular orienta- tion at the surface could be interpreted from irreversible potential-dependent behavior of the surface Raman spectrum. The structure of the reduction product was determined with confidence. Similar studies can be applied to other molecules. The new researcher-controlled experimental parameter of continuous electrode potential variations has direct application to systems incorporating electron transfer. The importance of electron transfer in biological systems makes those systems logical candidates for study by surface Raman techniques. Generally, Raman studies of large biologically important molecules have been concentrated in resonance Raman investigations of the chromophore. These studies are very effective in determining contributions to UV/vis absorption spectra by different chromophores and distinction 155 156 of similar chromophores in regard to substituents which may couple to the electronic transitions. The influence of the protein substructure on energy transfer is often invoked to describe dynamics observed in the chromophoric resonance Raman spectrum. The actual structure of many proteins has eluded determination owing to the inability to crystallize many of the large molecular weight molecules. The dynamics of protein structure change during oxidation and reduction at the redox center is essentially unknown. Because of the intense light scattering observed from molecules adjacent to the metal electrode surface in the SERS experiment it is reasonable to assume that the outer amino acid residues of an adsorbed protein should exhibit intense SERS signals. Thus, the SERS experiment may be applicable to analysis of structure, amino acid sequence, and energy transfer at the outer sections of a large protein molecule. The combination of potential dependent Spectroscopic phenomena and structure and orientation interpretations with excitation wavelength dependence of an adsorbed protein promises to make deep inroads into the Imechanisms of energy transfer in biologically important molecules. Of course, distinction of amino acid residues and their orientation at the electrode surface are pre- requisites to any interpretation of outer protein structure. In this chapter the surface enhanced Raman spectrum of two small amino acids, L-glycine (NHZCHZCOO-) and L-leucine 157 [(CH3)2CHCHNH COO-l are presented and interpreted in terms 2 of orientation at the electrode surface. It is hoped that distinctions in the surface enhanced Raman spectrum of these and other amino acids may be extrapolated to assign— ments in surface spectra of adsorbed protein molecules in the future. The intensity of the surface enhanced Raman signals from the adsrobed amino acids are significantly reduced from the signals observed from t-l,2-bis(4-pyridyl)ethylene on silver. The signals observed from the amino acids on silver generally were in the range 3 x 103 counts per second, or less, whereas the signals obtained from t-BPE on 5 to 3 X 105 counts per second. The silver ranged from 1 X 10 reasons for the sharp drop in intensity are consistent with the ligating ability of t-BPE, over that of the amino acids, in conjunction with a resonance Raman enhancement mechanism involving a silver complex. The normal Raman scattering cross section of the highly conjugated t-BPE molecule over that of the saturated amino acids examined here may also account for the remarkable intensity of t-BPE surface Raman signals. As mentioned previously, the hydrophobic character of t-BPE undoubtedly assists in the facile adsorptions of the bipyridine to the silver surface. The amino acids are polar molecules and are therefore hydrophi- lic. Notwithstanding the low intensity of the surface Raman signals from glycine and leucine some interpretations of 158 molecular orientation at the electrode surface can still be made. The Raman spectra of L-leucine and L-glycine on silver at -600 mV under 5145 A excitation are shown in Figure 22 and the frequencies and assignments are listed in Table 9. These spectra were recorded in a bulk pH of 11.0. Since at this pH both amino acids are above their respective isoelectric points the molecules carry a nega- tive charge associated withtflmacarboxylate group. No surface Raman spectrum was observable from amino acids at the silver electrode when the pH of the bulk solution was below the isoelectric point of the amino acids. Therefore, it can be assumed that the negatively charged amino acids adsorb more strongly to the positively polarized electrode than do the neutral compounds in agreement with the t-BPE results. The pH adjustment from neutral solution wasaccomplished by addition of KOH. The formation of silver oxide (A920) and silver hydroxide (AgOH) was possible in the hydroxide media. In order to determine the presence of the oxides at the surface and their influence on the surface Raman spec- trum observed for the amino acid system the normal Raman spectra of the two polycrystalline oxides was obtained. Both A920 and AgOH are black opaque solids which required a back scattering geometry for observation of the normal Raman spectra. These spectra are shown in Figure 23 and the frequencies listed in Table 10. The normal Raman spectra of these two silver oxides have not been previously 159 Figure 22: Surface Raman Spectra of L-leucine and L-glycine on Silver at -600 mV (sce). + EXCITATION WAVELENGTH: 5145 A (Ar ). p—_ _ . 5145 I Amino Acids on Silver at -600 mv VS. SCE. Excitation LLEI l- leucine 1051 £09 119 £81 900 0Z6 0001 1201 9101 0911 £111 891! 0181 OZ?! 9771 £15! 6851 1291 £59! 160 l- Glycine uogssgwa lugseI-uou 600 1400 1600 [IgsuaIuI 161 TABLE 9: Raman Spectra of L—Glycine and L-Leucine adsorbed onto a polycrystalline silver electrode at -600 mV sce, in the frequency region 550-1700 cm’l. Excitation: 5145 A. _——_—*_____— _...-. - _.___.___.. Assignment L-glycine L-leucine NH2 torsion 555 557 C00” wag 605 607 coo" bend 671 teflon? 736 790 783 806 CH2 rock 920 CH3 rock 1000 CCN asymmetry stretch 1035 1027 w 1078 1076 NH2 rock 1142 1140 stretch of CH-(CH3)2 1177 CH2 twist 1270 1268 CH2 wag 1310 C00' symmetry stretch 1380 1377 methyl group motion? 1424 CH2 bend 1450 1446 NH2 symmetry deformation 1498 1501 1577 C00' asymmetric stretch 1588 1589 NHZ degenerate deformation 1621 1653 162 Figure 23: Normal Raman Spectra of Polycrystalline A920 and AgOH Utilizing a Back-scattering Geometry. EXCITATION WAVELENGTHS: 5145 A (Ar+) AgOH 6471 A (Kr+) A920 I.ma 0001 0091 0071 0021 009 009 002 163 INTENSITY L- -— 1560 1510 I H- Q? h (:9 £r: 1300 = c l ‘l non-lasing emission I l 1050 979 957 104 1 115 678 see 538 “E 492 he 3 403 332 348 non-lasing emission 290 280‘,69 232 1489 butane: Suglaueos noes 164 TABLE 10: Raman Spectra of Polycrystalline AgOH and AgZO in the region 200-1600 cm- using a back scattering geometry. —_. __—__._._.—_-- w —.———-—. .. _—...-..- . .— Polycrystalline Polycrystalline AgOH AgZO 232 269 290 286 348 382 403 492 538 588 678 715 704 729 748 857 957 970 979 980 1050 1489 1300 1560 1510 165 :reported in the literature though the infrared spectrum (of A920 has been reported.112"113 No attempt at assignment of observed frequencies is made. Further study is required for such assignment. Qualitatively, the effect of hydrogen loonding in the back scattering spectrum of AgOH is mani- zfested in the broadening of all Raman lines in comparison ‘to those observed in the spectrum of AgZO. The spectra of 'the two silver compounds were obtained mainly to examine ‘their interference in the surface Raman spectrum of the amino acids. No interference was observed, however, and it can be concluded that either the silver oxides are not present or that the intensity of their surface signal is too weak to be observed. Returning now to the surface Raman spectra of the amino acids shown in Figure 22 some interesting anomalies are observed. Assignments of the observed surface Raman spectra for L-leucine and L-glycine are made by comparison to other vibrational spectroscopic studies of amino acids.ll4-120 The suggested assignments are listed in Table 9. The most important and distinct difference between the surface spectra of L-leucine and that of L-glycine is the appearance of a line at 671 cm.1 in the leucine spectrum which is absent from the glycine. This mode has been assigned to the ~COO- bend in leucine and is expected to occur in the same region of the glycine surface spectrum. The -COO_ bending vibration occurs at 690 cm"1 in the infrared spectrum of glycine and at 694 in the Raman spectrum 166 of polycrystalline y-glycine.117 The most intense line in both surface spectra appears at 1377 cm-1 in L-leucine and 1380 cm-1 in L-glycine and is probably the symmetric stretching motion of the -C00- group. The line at 1424 cm-1 in the surface spectrum of L-leucine has no counterpart in the spectrum of L-glycine and is accordingly assigned to motion in the methyl substituents in leucine. The COO- asymmetric stretch appears as a doublet at 1589 cm-1 and 1577 cm-1 in the surface spectrum of L- leucine, but as a singlet at 1588 cm.1 in the spectrum of L-glycine. The appearance of splitting in the asymmetric stretching motion of the C00. group in the leucine spectrum may be a consequence of bonding of one oxygen atom at the surface while the other remains free. This explanation is also invoked to account for the appearance of a C00- bend in the leucine spectrum and not in glycine. In glycine if both oxygen atoms are bound to the surface the bending vibration would be significantly dampled and probably shifted in frequency. The -NH2 deformations above 1600 cm-1 in the surface Raman spectrum of L-leucine appear more intense relative to the same vibrations in the glycine spectrum. This may also be a manifestation of the way in which glycine is attached to the surface. If the glycine molecule is attached through both carboxylate oxygens it may be forced into a configuration which keeps the -NH2 group from coming in contact with the surface. With only one oxygen attached in the leucine situation the molecules 167 maintain flexibility on the surface and allow the amine group to contact the surface. The clear distinction between the surface Raman spectra of L-leucine and L- glycine on silver may be extrapolated to distinguish the same amino acid residues in a protein adsorbed at an electrode surface. Future Surface Raman Studies. The analysis of experimentally observed surface Raman spectra presented in this thesis can be applied to nearly any molecule. Molecules which have demonstrated ability as ligands in metal complexes undoubtedly will exhibit intense surface Raman scattering. Those molecules which are of particular interest to the verification of the surface dynamics of t-l,2-bis(4-pyridyl)ethy1ene on silver are the isomers to the highly symmetrical 4,4'- bipyridylethylene (i.e. the isomers with nitrogens in positions other than the 4 position). The very low frequency region (0-100 cm-l) of the surface Raman spectrum of t-BPE should demonstrate significant spectroscopic changes accompanying the molecular reorientation from the end-on to the flat configuration. Gerack et a1.121 attributed potential dependent dynamics of the low frequency surface spectrum of pyridine to reorientation of the molecule to the "standing up" position at less negative potentials. Determination of surface coverage is particularly important to the designation ofenlenhancement factor. The 168 monoprotonated t-BPE molecule generates a surface Raman spectrum distinct from the unprotonated molecule and may therefore be exploited in accurate determination of surface coverage through titration. The titration tech- nique has some obvious advantages over electrochemical techniques which require sweeping of electrode potential. Also the surface Raman spectrum of monoprotonated t-BPE is observed even at very light coverages (250 nM bulk concentrations). Electrochemical measurements are gener- ally insensitive or irreproducible at such low coverages. Excitation profiles in regions of the visible spectrum other than the rhodamine 6G region are necessary to analyze the energy distribution of eigenstates in t-BPE and other adsorbates. Profiles should also be obtained at other potentials, as shown to be important in the discussion of potential-dependent surface Raman intensities. The use of teflon as an internal standard should be exploited in the preparation of the excitation profiles in that C104- demonstrates its own potential dependent behavior at the electrode surface. The application of SERS to biological molecules should be accelerated. The advantages of observation of these molecules at the electrode surface include flour- escence quenching by the metal surface, absence of photodecomposition problems owing to the behavior of the metal as a heat sink, and the ability to vary continuously 169 potential while observing the effect of electron transfer on the surface Raman spectrum. Further experimentation and theoretical investigation on the influence of the multiple enhancement mechanisms, proposed to account for the anomalous surface Raman intensities, are necessary, but at this time applications of the SERS technique are more generally worthwhile. The most important conclusion which can be drawn from this thesis is that any discussion of origination of Raman intensities from surface adsorbates must include mention of surface complex formation and the resonance Raman intensity enhancement expected from the perturbation of molecular electronic state distributions. 170 APPENDIX A The form of the normal coordinates of some pyridine vibrational modes, as determined by Long and Thomas,107 are depicted in this appendix. Discussion of the form of the normal coordinates was central to the assignment of lines in the normal Raman spectrum of t-l,2-bis(4- pyridyl)ethylene (see Chapter 5). 171 List of References 172 10. 11. 12. 13. 14. 15. 16. 173 REFERENCES T. Katsumoto; Bull. Chem. Soc. Japan 33, 1376 (1960). . M. Fleischmann, P.J. Hendra, A.J. McQuillan; Chem. Phys. Lett. 23(2), 163 (1974). R.L. Paul, A.J. McQuillan, P.J. Hendra,14.Fleischmann; J. Electroanal. Chem. 33, 245 (1975). A.J. McQuillan, P.J. Hendra, M. Fleischmann; J. Electroanal. Chem. 33, 933 (1975). D.L. Jeanmaire, R.P. van Duyne; J. Electroanal. Chem. 33, l (1977). D.L. Jeanmaire, R.P. van Duyne; J. Electroanal. Chem. 33, 235 (1975). J.A. Creighton, M.G. Albrecht, R.E. Hester, J.A.D. Matthew; Chem. Phys. Lett. 33, 55 (1978). B. Pettinger, U. Wenning, D.M. Kolb; Ber. Bunsenges. Phys. Chem. 32, 1326 (1978). J.A. Creighton, C.G. Blatchford, M.G. Albrecht; J. Chem. Soc. Faraday II 13, 790 (1979). M. Moskovits; J. Chem. Phys. 33, 4159 (1978). M. Moskovits; Solid State Comm. 32, 59 (1979). J.A. Creighton, C.G. Blotchford, J.R. Campbell; Proc. of the VIIth Int. Conf. on Raman Spectroscopy, 398-399 (1980) (W.G. Murphy, ed.). J.I. Gersten; J. Chem. Phys. 22(10), 5779 (1980). F. King, G.C. Schatz, and R.P. van Duyne; J. Chem. Phys. 33, 4472 (1979). G.C. Schatz, R.P. van Duyne; Surface Sci. 101, 425 (1980). C.S. Allen, G.C. Schatz, R.P. van Duyne; Chem. Phys. Lett. 13(2), 2101 (1980). 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 174 S. Efrima, H. Metiu; Chem. Phys. Lett. 99(1), 59 (1978). S. Efrima, H. Metiu; J. Chem. Phys. 29(4), 1602 (1979). S. Efrima, H. Metiu; J. Chem. Phys. 29(5), 2297 (1979). S. Efrima, H. Metiu; J. Chem. Phys. 29(4), 1939 (1979). S. Efrima, H. Metiu; Israel Journal of Chem. 29, 17 (1979). G. Karzeniewski, T. Maniv, H. Metiu; Chem. Phys. Lett. 29(2), 301 (1980). P.K. Avavind, H. Metiu; Chem. Phys. Lett. 29(2), 301 (1980). M.R. Philpott, J. Chem. Phys. 92(5), 1812 (1975). E.N. Economon, K.L. Ngai; Adv. Chem. Phys. 22, 265 (1974). Y.Y. Teng, E.A. Stein; Phys. Rev. Lett. 29, 511 (1967). D. Beaglehole; Phys. Rev. Lett. 22, 708 (1969). I. Pockrand; J. Phys. 99, 2423 (1976). . Raether, Nuovo Cimento 299, 817 (1977). Raether, Phys. Thin Films 9, 145 (1977). H H A. Girlando, M.R. Philpott, D. Heitmann, J.D. Swalm, R. Santo, J. Chem. Phys. 22(9), 5187 (1980). A H A . Girlando, J.G. Gordon II, D. Heitmann, M.R. Philpott, . Seki, J.D. Swolen; Surf. Sci. 101, 417 (1980). . Otto; Z. Phyzik 241, 398 (1968). E. Kretschmann; Z. Phyzik 241, 313 (1971). R.M. Hexter, M.G. Albrecht; Spectrochim. Acta 35A, 233 (1979). A. Brillante, I. Pockrand, M.R. Philpott, J.D. Swalen; Chem. Phys. Lett. 92(3) 395 (1978). 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 175 W. Carius, O. Schroter; Experimentelle Technik der Physik 29(1), 35 (1980). B. Pettinger, A. Tachjeddime, D.M. Kolb; Chem. Phys. Lett. 99(3), 544 (1979). W.P. Chen, J.M. Chen; Surf. Sci. 92, 601 (1980). B. Pettinger, U. Wenning, H. Wetzel; Surf. Sci. 101, 490 (1980). R. Donnhaus, R.E. Benner, R.K. Chang, I. Chabay; Surf. Sci. 101, 367 (1980). A. Otto; Surf. Sci. 101, 99 (1980). I. Pockrand, A. Brillante, D. Mobins; Chem. Phys. Lett. 99(3), 499 (1980). J.G. Gordon II, S. Ernest; Surf. Sci. 101, 499 (1980). A. Regis, J. Corset; Chem. Phys. Lett. 29(2) 305 (1980). A. Regis, J. Corset; Extended Abstracts, 81—1 Spring Meeting of the Electrochemical Society, 921 (1981). M.W. Howard, R.P. Cooney, A.J. McQuillan; J. Raman Spec. 9(4), 273 (1980). R.P. Cooney, A.J. McQuillan, M.W. Howard; Proc. of the VIIth Int. Conf. on Raman Spectroscopy, 400 (1980) (W.G. Murphy, ed.)T J.C. Tsang, J.E. Demuth, P.N. Sanda, J.R. Kirtzey; Chem. Phys. Lett. 29(1) 54 (1980). G.W. Robinson; Chem. Phys. Lett. 29(2), 191 (1980). E. Burstein, Y.J. Chen, C.Y. Chen, S. Lundquist, E. Tosatti; Solid State Comm. 29, 567 (1979). C.Y. Chen, E. Burstein, S. Lundquist; Solid State Comm. 92, 63 (1979). R.R. Smardzewski, R.J. Colton, J.S. Murday; Chem. Phys. Lett. 99(1), 53 (1979). S.R. Kelemen, A. Kaldor; Chem. Phys. Lett. 29(2), 205 (1980). 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 176 J.E. Demuth, K. Christmann, P.N. Sanda; Chem. Phys. Lett. 29(2), 201 (1980). R.E. Chance, A. Prock, R. Silbey; Adv. Chem. Phys. 92, 1 (1978). D.A. Zwemer, C.V. Shank, J.E. Rowe; Chem. Phys. Lett. 29,(2), 201 (1980). J.E. Rowe, C.V. Shank, D.A. Zwemer, C.A. Murray; Phys. Rev. Lett. 99(26), 1770 (1980). H. Seki, M.R. Philpott; J. Chem. Phys. 29(10), 5376 (1980). C.Y. Chen, I. Davoli, G. Ritchie, E. Burstein; Surf. Sci. 101, 363 (1980). J.P. Heritage, J.G. Bergman, A. Pinczuk, J.M. Worlock; Chem. Phys. Lett. 92(2,3), 229 (1979). R. Dornhaus, M.B. Long, R.E. Benner, R.K. Chang, Surf. Sci. 99, 240 (1980). G.R. Erdheim, R.L. Birke, J.R. Lombardi; Chem. Phys. Lett. 99(3), 495 (1980). W. Krasser, H. Ervens, A. Fadini, A.J. Renouprez; J. Raman Spec. 9(2) 80 (1980). M. Moskovits, D.P. Dilello; J. Chem. Phys. 29(12), 6068 (1980). T.M. Cotton, S.G. Schultz, R.P. van Duyne; J. Am. Chem. Soc. 102, 7960 (1980). M.B. Lippitsch; Chem. Phys. Lett. 29(2), 224 (1981). E. Koglin, J.M. Sequaris, P. Valenta; J. Mol. Str. 99, 421 (1980). M. Moskovits, D.P. DiLella; Chem. Phys. Lett. 29(3), 500 (1980). M.R. Philpott, J.G. Gordon II, B. Pettinger, Extended Abstracts, 81-1 Spring Meeting of the Electrochemical Society, 896 71981). T.H. Wood, M.V. Klein; J. Vac. Sci. & Tech. 29, 459 (1979). M. Moskovits, D. DiLe11a, P. McBreen, R. Lipson, A. Gohin; Proc. of the VIIth Int. Conf. on Raman Spectroscopy, 394 (1980) (W.G. Murphy, ed.). 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 177 H. Wetzel, B. Pettinger, U. Wenning; Chem. Phys. Lett. 29(1), 173 (1980). E. Burstein, C.Y. Chen, S. Lundquist; Proc. of the Joint US-USSR Symposium on Theory of Light Scattering in Condensed Matter, (Plenum Press, New YOrk, 1979). R.P. van Duyne; Chemical and Biochemical Applications of Lasers, Vol. 4, (C.B. Moore, ed), (Academic Press, New York, 1978). T.E. Furtak, Solid State Comm. 29, 903 (1978). T.E. Furtak, G. Trott, B.H. Loo; Surf. Sci. 101, 374 (1980). W.A. Goddard III, T.C. McGill; J. Vac. Sci. Technol. 29(5) 1308 (1979). A. Otto, J. Timper, J. Billman, G. Kovacs, I. Pockrand; Surf. Sci. 92, L55-L57 (1980). A. Otto; Surf. Sci. 92, 145 (1980). J.I. Gersten; J. Chem.Phys. 22(10), 5779 (1980). H. Wetzel, H. Gerischer, B. Pettinger; Chem. Phys. Lett. 29(2), 392 (1981). J.R. Lombardi, E.A. Shields Knight, R.L. Birke; Chem. Phys. Lett. 29(2), 214 (1981). J. Billmann, A. Otto, I. Pockrand, J. Timper; Proc. of the VIIth Int. Conf. on Raman Spectroscopy, 424 (1980) (W.G. Murphy, ed.). B. Pettinger, H. Wetzel; Chem. Phys. Lett. 29(2), 398 (1981). J.R. Lombardi, R.L. Birke; Extended Abstracts 81-1 Spring Meeting of the Electrochemical Society, 894 (1981). D.L. Rosseau, J.M. Friedmann, P.F. Williams, in: Topics in Current Physics, Vol. 2 Raman Spectroscopy of Gases and Liquids, (A. Weber, ed.), chapter 6. D.K. Lavalee, E.B. Fleischer; J. Am. Chem. Soc. 99(8), 2583 (1972). H.B. Gray, C.J. Ballhausen; J. Am. Chem. Soc. 99, 260 (1963). 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 102a. 103. 104. 105. 178 H.A. Pearce and N. Sheppard; Surf. Sci. 99, 205 (1976). R.M. Hexter, M.G. Albrecht, Spectrochim. Acta 35A, 233 (1979). A.C. Albrecht, J. Chem. Phys. 99(5), 1476 (1961). J. Tang, A.C. Albrecht, in Raman Spectroscopy, Theopy and Practice, Vol. 2, pp. 33-68 (1970) (H.A. Szymanski, ed.). R.J.H. Clark, B. Stewart; Structure and Bonding 99, 1 (1979). S.L. Ross; Introduction to Ordinary Differential Equations (Xerox College Publishing, MA) (1966) Chapter 5, pp. 152-160. B. Pettinger, U. Wenning, H. Wetzel; Chem. Phys. Lett. 92(1), 192 (1979). U. Wenning, B. Pettinger, H. Wetzel; Chem. Phys. Lett. 29(1), 49 (1980). D.P. Shoemaker, C.W. Garland, J.I. Steinfeld; Experi- ments in Physical Chemistry, 3rd ed. (1974) pp. 455- 457. Chemical Rubber Co; Handbook of Chemistry and Physics, 52nd ed. p. B232. H.H. Perkampus, E. Baumgarten; Spectrochim. Acta 29, 359 (1964). E.B. Wilson; Phys. Rev. 99, 706 (1934). E. Spinner; J. Chem. Soc., 3870 (1963). G. Valette, A. Hamelin; J. Electroanal. Chem. 99, 301 (1973). G. Herzberg; Molecular Spectra and Molecular Structure II Infrared and Raman Spectra of Polyatomic Molecules, (Van Nostrand Reinhold Co.) (1945), 247-249. F.R. Dallish, W.G. Fateley, P.F. Bentley; Characteris- tic Raman Frequencies of Organic Compounds (Wiley- Interscience) (1974) pp. 263-272. Z. Meic, H. Gusten; Spectrochim. Acta 34A, 101 (1978). 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 179 A. Regis, J. Corset; Extended Abstracts 81-1, Spring Meeting of the Electrochemical Society, 921 (1981). D.A. Long, E.L. Thomas; Trans. Faraday Soc. 99, 783 (1963). E. Spinner; Aust. J. Chem. 29, 1805 (1967). R.H. Dyck, D.S. McClure; J. Chem. Phys. 99(9), 2326 (1962). J. Volke, J. Holubek; Coll. Czech. Chem. Commun. 22, 1777 (1962). W.L. Jorgensen, L. Salem; The Organic Chemists Book of Orbitals (Academic Press, NY) (1973). E.F. Gross, F.I. Kreingold; Optics and Spectroscopy 29, 211 (1961). V.M. Kirillova, E.N. Lotkova; Optics and Spectroscopy 29, 181 (1966). M. Tsuboi, T. Onishi, I. Nakagawa, T. Shimanouchi, S. Migushima; Spectrochim. Acta 22, 253 (1958). R.K. Khanna, P.J. Miller; Spectrochim. Acta 26A, 1667 (1970). R.K. Khanna, M. Horak, E.R. Lippincott; Spectrochim. Acta 22, 1759 (1966). M.A. Salimov, V.A. Pchelin, A.V. Kerimbekov; Russ. J. Phys. Chem. 92(10), 1231 (1963). M. Tsuboi, T. Takenishi, A. Nakamura; Spectrochim. Acta 29, 271 (1963). K. Machida, A. Kagayama, Y. Saito, Y. Kuroda, Y. Uno; Spectrochim. Acta 33A, 569 (1977). R.S. Krishnan, V.N. Sankaranarayanan, K. Krishnan; J. Ind. Inst. Sci. 99(2), 66 (1973). A.Z. Genack, D.A. Wertz, T.J. Gramila; Surf. Sci. 101, 381 (1980). MICHIGAN STATE UNIVERSITY LIBRARIES IIIllUllllHI("WNWIIHIIIIUIHIIHIIIHIUIIPIIIHIM 3 1293 03145 5805