LIBRARY Michigan State University This is to certify that the thesis entitled Spectrophotometric, Electrochemical, Photoelectrochemical, and Surface Analysis Studies of Copper Phthalocyanine/Metal Oxide Electrodes presented by Vance Roger Shepard, Jr. has been accepted towards fulfillment V of the requirements for M . S . degree in Chemi $12er , k “s - “Ru“!!flfi Major professor ‘ Date \\\\\‘\'\ \ \ 0-7 639 SPECTROPHOTOMETRIC, ELECTROCHEMICAL, PHOTOE ECTROCHEMICAL, AND SURFACE ANALYSIS STUDIES OF COPPER PHTWALOCYANINE/METAL OXIDE ELECTRODES 8V Vance Rogers Shepard, Jr. A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Chemistry 1977 ABSTRACT SPECTROPHOTCMETRIC, ELECTROCHEMICAL, PHOTOELECTROCHEMICAL, AND SURFACE ANALYSIS STUDIES OF COPPER PHTHALOCYANINE/METAL OXIDE ELECTRODES By Vance Rogers Shepard,Jr. There has been a growing interest in the study of chemically-modified electrode surfaces. Irreversible adsorption or covalent-attachment of various molecules to an electrode surface can impart specific catalytic prop- erties which the electrode alone did not possess. Copper phthalocyanine (CuPc) in its solution form, strongly ad- sorbed, or covalently-attached to a metal oxide electrode (SnO or Ti02) was studied. Spectrophotometric methods 2 aided in determination of the number of adsorbed or co- valently-attached dye molecules present on the electrode surface. Differential capacitance measurements for SnO2 and TiO and cyclic voltammetry of the tetrasodium salt of 2 tetrasulfonated copper phthalocyanine (CuPc(SOSNa+)h) in DMSO and H20 resulted in energy mapping of the semiconductor band structure in relation to the redox couples of the dye. Cyclic voltammetry of adsorbed and covalently-attached dye indicated retention of an electrochemically active species on the surface. ESCA analysis of electrodes in various states of electrochemical treatment showed two types of phthalocyanine present and variation in the copper valence state with change of solvent. Photocurrents were generated by the adsorbed and covalently-attached CuPc electrodes. Photocurrent studies indicated a potential dependent decom- position/desorption process of the adsorbed dye electrodes. A steady potential dependent response was observed for the covalently-attached CuPc electrodes. Dedication To Mom and Dad i1 "ACKNOWLEDGMENTS" I would like to thank Neil and Monty for their advice and encouragement. I would like to thank my parents, grand- parents, Mrs. Elmer Clark, Mr. Alden Eddy and friends for their encourgement, faith, and thoughts. And many thanks to Victoria for her help in the preparation of this thesis. iii CHAPTER I. A. B. C. D. E. CHAPTER II A. B. F. G. CHAPTER III. A. "TABLE OF CONTENTS" INTRODUCTION. . . . . . . . . . Semiconductors (n-Type). . . . . . Chemical Modification of Electrode ESCA . . . . . . . . . . . . . . . Photocurrent Response Studies. . . Phthalocyanines. . . . . . . . . . . EXPERIMENTAL . . . . . . . . . Dyes .-. . . . . . . . . . . . . . Solvents and Electrolytes. . . . . Spectrophotometric Studies . . . . Electrochemical Studies. . . . . . Covalent-Attachment. . . . . . . . 1. Sulfonamide Attachment . . . 2. Thiol. . . . . . . . . . . . Photocurrent Studies . . . . . . . ESCA Analysis. . . . . . . . . . . RESULTS AND DISCUSSION. . . . Solution Studies . . . . . . . . . 1. Spectrophotometric Studies of CuPc(SO O 3 Na+)h. . . . . . . . . . . . . . . . 2. Electrochemistry of CuPc(SOSNa+)h. . . iv 10 16 17 18 18 19 20 21 21 22 22 25 26 26 . 26 B. Adsorbed Dye Studies. . . . . . . . . . 1. Spectrophotometric Studies of CuPC(SOSN8+)4 . . . . . . . . . 2. Electrochemistry of CuPc on Snoz. 3. ESCA of CuPc on Sn02. . . . . . . 4. Photocurrent Response of CuPc on 31102 C. Covalent-Attachment Studies . . . . . . 1. Spectrophotometric Studies of Covalently-Attached CuPc. . . . 2. Electrochemistry of Covalently- Attached CuPc-Sn02 Electrodes . 3. Photocurrent Response . . . . . . CHAPTER IV. SUGGESTIONS FOR FUTURE WORK . . . . LIST OF REFERENCES. 0 O C O O O O O O O O O O O C 41 41 43 45 52 58 58 60 64 67 69 "LIST OF TABLES" Page Table 1. ESCA Data CuPc/Sn02. . . . . . . . . . . . . .49 vi Figure Figure Figure 3 Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure 13. 1h. "LIST OF FIGURES" Energy diagram of semiconductor-electrolyte interface. . . . . . . . . . . . . . . . . 4 Structure of phthalocyanine. . . . . . . . . 11 CuPc covalent-attachment schemes . . . . . . 15 Absorption spectra of CuPc(SOSNa+)LL in DMSC and H20. 0 e e o e o o o e e e o e o e o o 27 Cyclic voltammetry of CuPc(SO§Na+)u in 0.1fiL TEAP/DMSO at (a) Sn02, (b) Ti02, and (c) Pt 0 O O O O O O O O I O O O O O O O O O O 30 Cyclic voltammetry of CuPc(SOSNa+)u in R20 (pH u) at (a) Sno2 and (b) TiO . 32 2. O O O 0 Differential capacitance vs potential plots of SnO2 in 0.1IL TEAP/DMSO and H20(pH h) . 35 Differential capacitance vs potential plots of TiO2 in 0.1ll TEAP/DMSO and R20(pH h) . 37 Energy diagram of CuPc(SOSNa+)u redox couples in 0.1 u TEAP/DMSO at SnOZ, Ti02and Pt . . 39 Energy diagram of CuPc(SOSNa+)LL oxidation in R20(pH h) at SnO2 and TiO2 . . . . . . . . 40 Absorption spectra of clean SnO2 and CuPc sub- limed on SnO2 electrodes . . . . . . . . . 42 Cyclic voltammetry of CuPc sublimed on SnO2 in0.1f.LTEAP/DMSO............ 44 Cu(2p1/2 3/2) ESCA spectra of. . . . . . . . 48 N(1s) ESCA spectra of. . . . . . . . . . . . 51 vii Figure Figure Figure Figure Figure 15. 16. 17. 18. 19. Photocurrent response of a CuPc/SnO2 elec- trode in 0.05 Na2CZOu/pH 7 buffer. . . . . 53 Photocurrent response vs anodic bias poten- tial plots of two CuPc/SnOP electrodes in 0.05M NaZCZOu/pH 7 buffer. . . . . . . . . 56 Absorption spectra of clean SnO2 and coval- ently-attached-SnO2 electrodes . . . . . . 59 Cyclic voltammetry of a covalently-attached CuPc-SnO2 electrode in 0.111 TEAP/DMSO . . 62 Photocurrent reaponse vs anodic bias poten- tial of a covalently-attached CuPc-SnO2 electrode. . . . . . . . . . . . . . . . .66 Viii CHAPTER I INTRODUCTION 2 The photodecomposition of water into its elements (H2, 02) is one of the most attractive means of storing solar energy. This decomposition process may be broken down into three steps: light absorption, water oxidation, and water reduction. Metal oxides (SnO2 and Ti02) doped as n—type semiconductors have catalyzed the decomposition of H20 (1 ). Certain dyes, by virtue of their high absorption in the visible region and their facility in undergoing oxidation-reduction reactions, should also be capable of mediating some or all of these processes. Dye in solution ( 2), adsorbed (.3), and covalently-attached (li,ES) at metal or semiconductor electrodes have been used to gener- ate photocurrents. A brief discussion of semiconductor electrochemistry, covalent-attachment (chemical modificat- ion), and photocurrents is essential for a better under- standing of light energy conversion. A useful tool for surface analysis, electron spectroscopy for chemical analysis (ESCA), is also discussed. Semiconductors(n-Type) Great progress in semiconductor electrochemistry has been made since realization of their potential use in solar cells. Various n-type semiconductors such as Sn02 and TiO2 have been used extensively for studies in both aqueous and nonaqueous media (6:). Electrode processes on semicon- ductors show certain characteristics that are different from those of electrode reactions on metals. The main 3 difference in reactions on metals and on semiconcuctors is that in the latter, the kinetics of the reaction may depend on processes that occur within the solid electrode. A bending of the bands occurs at the surface of a semiconductor in contact with an electrolyte. Application of an anodic potential results in formation of a depletion or positive charged layer at the semiconductor surface (or in the space charge region within it), thus leading to upward bending of the bands (Figure 1). Application of a cathodic potential gives rise to a downward bending of the bands due to formation of a negative charge layer (7 ). When a semiconductor is in contact with an electrolyte, a significant part of the potential drop is across the space charge region within the semiconductor; the rest is across the double layer on the electrolyte side. In experimental investigations of the double layer at semi- conductor-electrolyte interfaces, an important parameter is the capacitance of the interface, C, which varies as a function of electrode potential V. The measured capacit- ance C is composed of 03c, capacitance of space charges in the semiconductor, and CH' the capacitance of the Helmholtz double layer on the solution side of the interface. Since the capacitances are in series and Csc is usually small as compared to C the measured capacitance (C) is essentially H' equal to Csc ( 7). A differential capacitance plot of 1/033 V8 potential according to the Mott-Schottky equation E.LL ./ Figure 1. Energy diagram of semiconductor-electrolyte interface. 5 1/C:c= 2/660eonOCDS- kT/eo yields information about the effective carrier density no from the slope of the plot and flatband potential vfb from the potential axis intercept. The flatband potential is the potential at which excess electrical charge in the semiconductor is zero and is indicative of the position of the Fermi level in the semiconductor band gap region with respect to the reference electrode. CSc is the capacitance of the semiconductor electrode, E and 60 are the dielect- ric constant of the semiconductor and permittivity of free space, respectively, so is the absolute value of the elect- ronic charge, no is the donor density, k is the Boltzmann constant, and.qb8, is the potential difference between the flatband potential and the potential at which the measure- ment is made (7 ,8 ). Chemical Modification of Electrode Surfaces Chemical modification of surfaces is useful in research areas such as chromatography and catalysts. Chemically- modified solid supports have been use to improve chromato- graphic column performance. This has depended on an organosilane reaction with the surfaces of silica or alumina particles. Homogeneous catalysts have been attached to silica to provide heterogeneous hydroformylation catalysts (<9). Acid-base dye indicators have been attached to silica surfaces to create solid indicators (10). Electrochemistry is a field where chemically-modified 6 surfaces are of great value. Chemical modification of an electrode surface refers to strong binding of a selected chemical reagent to the surface to endow it with some or all of the chemical and electrochemical properties of the selected reagent. Such chemically-modified electrodes rep- resent new approaches to the study of electrochemical reactions. Modification based on covalent bond formation between the reagent and electrode has been described for carbon (11), Sno2 (12,13), and T10 (1h). Optically active amino acids 2 have been bound to carbon electrodes via amide bonds to form a chiral electrode (15). The work on SnO2 and TiO2 elect- rodes utilize organosilane reagents, which bear amine, pyridyl, mercapto, and other functional groups. Schematically the reaction of an organosilane with a surface hydroxyl group can be represented as a a -M OH + SlYZR --§ -M€§—-OSiYZR + HA (13). en face silane ‘ urface The modified surface chemical stability is quite good except in the presence of strong acid or base solution (13). Rhodamine B and iron porphyrin have been attached to SnO via an amide bond tit.16). Recently, Hawn and Armstrong 2 attached erythrosin to SnO2 using either an amide or a thiol linkage (53). These chemically-modified electrodes have displayed some very interesting electrochemical and photo- electrochemical results in which the electrodes appear to be chemically stable. ESCA The ESCA (electron spectroscOpy for chemical analysis) technique was originated and developed by Siegbahn and co- workers in Sweden during the 1960's (17,18). In ESCA, nearly monochromatic x-radiation of 125hev (Mg K11'2) energy impinge on a surface. Radiation in this range provides sufficient energy for the ejection of core electrons. Al- though the x-rays pentrate to thousands of angstroms, the photoelectrons produced have an escape mean free path of only S A0 to 100 A0, depending on the material. Core elect- rons kinetic energies, Ek’ are measured accurately, making it possible to determine the binding energies, Eb, of the core electrons from the relation Eb = hv I Ek ' qup’ where¢>8p is the work function or energy necessary to raise a free electron from the Fermi level to vacuum level . Binding energies are responsive to changes in the chemical environment. Shifts in binding energies are produced by the charge and valence state of an atom and the electronic relaxation effect from neighboring atoms. This refers to the effect of electronic charge flow from neighboring atoms to the core hole in the positive ion produced by the photo- emission (19,20). ESCA is sensitive to all the elements but hydrogen and helium. Since the measured electron binding energies of most elements are unique, a broad range scan of electron energy spectrum provides a good means of qualitative 8 analysis. The sensitivity of ESCA is on the order of 10"6 grams. The integrated intensities of the electron signals are directly proportional to the number of similar atoms in the sample, representing a pseudo-quantative tool. To insure minimum surface contamination, mainly carbon, a 10-8 to 10'10 torr vacuum is required in the sample chamber. Samples may be cleaned or depth profiled by an argon sputter gun (19). ESCA has been used extentfully in adsorption (21), surface oxide (22), and catalysts (23) studies. Surfaces of clean (12.2h) and chemically-modified (13,11) semiconductor electrodes have been characterized. This has lead to a better understanding of the chemical and physical structure of these surfaces. Ehotocurrent Response Studies n-Type semiconductors such as T102 and SnO2 will eat- alyze the oxidation of water when irradiated with greater than bandgap light energy (1 ,25). The electrooxidation of water occurs as a result of vacancies produced in the valence band of the highly-doped semiconductor (7 ). .A depletion or space charge layer is maintained at the electrode surface by means of an applied bias potential. Under these conditions, the conduction band remains charge depleted, and holes created in the valence band move to the surface and accept charge in the oxidation of H20 or OH'. It is well known that certain photoexcited dye molecules in the vicinitv of the electrode interface are capable of 9 transferring charge to the conduction band of an n-type semiconductor electrode (2, 26). Figure 1 explains this process in terms of an energy diagram with the redox react- ions of the dye in both the ground and excited states. Photoexcitation of the dye occurs at light energies less than the semiconductor band gap energy. The excited dye is than capable of losing an electron to the conduction band of the semiconductor. Anodic bias results in charge depletion of the conduct- ion band, and electron flow is from the dye towards the semi- conductor conduction band. Photocurrent response as a function of wavelength usually parallels the absorption spectrum of the dye (22). If a reducing agent such as sod- ium oxalate is present in solution with the dye, enhanced photocurrents are observed (supersensitization). This is due to a continual regeneration of the ground state of the dye by the reducing agent. The kinetics of the supersensi- tiaztion reaction depend upon the extent of overlap of the energy distribution of the excited dye molecule population, redox states of the reducing agent, and appropriate energy level distribution of the semiconductor (26). Photocurrent response studies have been carried out on dyes in solution (2 ,26 ), adsorbed ( 3), or covalently- attached (it) at both metal and semiconductor electrodes. All of the dyes absorb strongly in the visible region. Large extinction coefficients are necessary for maximum capture of the sensitizing radiation. In most cases, 1O multilayers of dye result in increased quenching and an ohmic resistance for electron transfer is formed. This prevents fast removal of electronic charge from the dye layer (2). Phthalocyanines Phthalocyanine is a large organic heterocycle contain- ing 7T electrons. Transition metal and post-transition metal ions can be coordinated within this ring (Figure 2) (27); these highly colored compounds are unique in many ways. They are insoluble in most organic solvents and only slightly soluble in solvents such as o-dichlorobenzene and pyridine. Various functional groups such as halogens, amines, sulfonic and carboxylic acids can be substituted for hydrogens on the four benzene rings (Figure 2) (27). These functional groups greatly affect the color and solubility of the phthalocyanines. Colors vary from greens to blues; this is important in the dye industry where the phthalocyanines have their greatest use. Absorption spectra of phthalocyanines in solution and sublimed on glass have been reported (28,29). They absorb strongly in the visible region, having large extinction co- 1 1 efficients (30,000M‘ cm‘ - 150,000M- cm'1). The strongest absorption is in the 600nm-750nm (2.1ev-1.8ev) region and corresponds toTT Trfi transitions of the ring system (30). Cyclic voltammetry studies of CuPc(SO§Na+)u in .1 TEAP/ DMSO at mercury show two reversible one-electron reductions n: H, SO3H' I, 80201, etc. m: cm“? Co+2, Ge+u, Si+u, etc. Figure 2. Structure of phthalocyanine. 12 and an oxdation (30). Chemical stability of the monoanion, dianion, and monocation is indicated by the reversible re- dox couples. The anodic redox couple of copper, zinc, and nickel phthalocyanine was inaccessible due to use of a mer- cury electrode. Absorption spectra of negative ions formed by sodium reduction of various phthalocyanines have indicated addition of electrons to the ring system and not the central atom (31). The ability for electrochemical reduction of oxygen by organic semiconductors such as phthalocyanines has attracted the attention of many researchers in connection with their possible use in fuel cells (32,33,3h ). Much of the phth- alocyanine electrochemistry reported is centered around this. Photoinduced reduction has been done on phthalocyanine films on carbon and platinum in various pH buffers. It has been shown that the catalytic properties of phthalocyanines are largely determined by the nature of the central atom (35). The order of decreasing electrocatalytic activity is Fe:>Co ;>Ni)>Cu;2H2(metal-free). This has indicated that phthalo- cyanines with higher magnetic moments (or paramagnetic sus- ceptibility) appear to exhibit electrocatalytic activity toward the oxygen reduction reaction (36). Recently, attachment of cobalt phthalocyanine to cross- linked polyacrylamide produced a stable oxidation catalyst with enhanced activity. It was coupled by means of cyanuric Chloride t0 N32 groups of the polymer matrix (37). ESCA analysis of phthalocyanines sublimed on COpper has 13 shown two types of nitrogen present in metal-free phthalo- cyanine (four equivalent central nitrogens and four equi- valent meso-bridging nitrogens). Presence of a metal atom in the ring equalizes the energy of these nitrogens and re- sults in a sharp nitrogen (18) peak with a weak satellite (38. 39 ). The choice of copper phthalocyanine for use in our studies was for the following reasons: 1) structural simlarity to naturally-occuring porphyrins, 2) large extinction coef- ficient, 3) ease of copper detection in surface analysis by ESCA. Our initial investigations of the CuPc in its solution form, strongly adsorbed, or covalently-attached to SnO2 or TiO2 electrodes are discussed. Spectrophotometric studies were used to determine regions of maximum absorption and ex- tinction coefficients for the CuPc solutions and electrodes. Absorbance values were used to determine the number of dye molecules present on the adsorbed or covalently-attached dye/SnO2 electrodes. Differential capactiance plots for SnO2 and TiO2 gave in- formation concerning the band structure: the Fermi level position and no, the donor density. Cyclic voltammetry of CuPc(SOSNa+)h in DMSO and H20 was done to determine the pos- itions of the redox couples at Sn02, T102, and Pt. From these, energy diagrams showing the relation between the semi- conductor band structure and redox couples of the dye could be drawn. CuPc sublimed on SnO2 was used in cyclic voltammetric 14 studies and the behavior compared to that of the solution form. ESCA analysis of electrodes in various states of electrochemical pretreatment in DMSO and H20 along with standards was done to aid in understanding their electro- chemical behavior. Covalent-attachment of CuPc to SnO2 and TiO2 was attempted via a sulfonamide or thiol formation using the sulfonyl chloride or tetraiodated form of the dye, respect- ively (Figure 3). The dye was coupled to the surface using various organosilanes which had either a terminal amine or mercapto functional group. Cyclic voltammetry of the covalent-attached dye-semi- conductor electrodes was carried out to determine the chemical stability and position of redox couples. The behavior of the covalently-attached dye was compared to that of the solution and adsorbed form. The spectrophotometric, electrochemical, and ESCA studies characterized the behavior of the dye in its various forms. This facilitated understanding of the photocurrent response data. Photocurrents of adsorbed and covalently- attached CuPc-Snoz electrodes in pH 7 buffer as a function of anodic bias potential were explored. A reducing or supersensitizing agent such as sodium oxalate was added to enhance the photocurrent response. 15 mosenom pcosnomppmupCoHa>oo omSU .m opswfim cm \Cm _ x Q n 0 $53.25)“. Oil) L??? _ + 1m...» .6 I 8.2. + cm 13% 2 0 cm 2: 2 / \/\ 2 / ~/ \I 2 Di I \l \ \ 40.1% Guam Owe ,cm I , i \ z z 2 .o~o z , .5 Guam azuriW/Q u m 01.6 QT. muolziaxofizmiox 4i \ fir\ m~o.u + :zifii 85. + 0\ z _ , HEW: / ll . Z 1 cm \cm p~o Gwen MO_ZS- kT/eO yields information about the effective carrier density nO from 32 38mvlsec (a) [EMA/cm2 [31.3 [ti-3uA/cm2 (b) "'—"'|500 mv Figure 6. Cyclic voltammetry of CuPc(SO3 at (a) Sn0 and (b) Ti0 Na+)u in H201pH u) 2 2° 33 the slope of the plot and the flatband potential Vfb from the potential axis intercept. The results of these measurements for SnO2 and TiO2 in DMSO and 820 are shown in Figures 7 and 8. The effective carrier densities, assuming dielectric con- stants of 12.5 and 127 for SnO2 and Ti02, respectively, were as follows: -1.65 1 0.1 volts vs AgRE (SnOZ/DMSO), -1.h3 t 0.1 volts vs Ag/AgCl (Sn02/H20), -2.0 1 0.1 volts vs AgRE (TiOZ/DMSO), and -1.h volts vs Ag/AgCl (Tiog/H2O)' From the donor density, flatband potentials and band gaps of SnO2 and TiO2 (3.5ev and 3.0ev, respectively) the energy positions of the valence band and conduction band edges (Ev and EC) are determined from the relation Ec(ev) = eOVfb + kT 1n (no/Nc where NO the density of states at the bottom of the conduc- tion band is about 10'19 carriers cm.3 in most semiconduc- tors (53). Figures 9 and 10 show the relation between the band struc- tures of the semiconductors and the redox couples of CuPc(SOS Na+)u in DMSO and H O. The reductions and oxidation of CuPc 2 (303Na+)u occur at potentials coincident with the band gap region of SnO2 and Ti02. The fact that the processes appear chemically reversible or pseudo-reversible in DMSO may indi- cate the contribution of surface states to the charge transfer process or electron tunneling through the narrow depletion region since the number of charge carriers is high. Surface states can be thought of as intermediate energy levels of narrouw width in the band gap region. Adsorbed species or sur- face defects of the semiconductor generate these surface states Figure 7. Differential capacitance vs potential plots of SnO2 in 0.1IL TEAP/DNSO and H2O(pR h). 55 xas_a>v .a._m =.~- 9—- +- C1! + 1 Guinea :34 3:... I l.) 1..- N b 4| :5 =5 .— b 1‘1 l fin J i. ___l IINO. L-e=. iTaa. ~e\_ lrss. 2. LVN—. ro- :1 i. o. r. 9. . u :. -- v-‘(nn .a / '. a a J. nJ'v 0".- p n ,l la / a; 1+- sf: ‘1’ 37 3...: :5 9... ‘1?" .1. «s .i q fir . q is. atmo. lessee NE. .4... sea 3:..- ts. ass? a: e 2 Na: .8. 38 Figure 9. Energy diagram of CuPc(SOSNa+)LL redox couples in O'1FL TRAP/DMSO at Sn02. TiO (Fc) at Pt is given for reference. 2 and Pt. Ferrocene 39 u a L umduflumauicamava m . w _ amdmadlsfl W m a l H 1.143: 5m: s m . a . w I 1, m m W >ma ashmfiu.fin- _ _ . _ _ i teammangen- n >oe - - _ asamswsew- . L madmasnaw- lungzu has? a a m sch? >00. u._ on _02co.om _OZC0.0& 7 _OZCO.O& xopea woven .xppea & «0: N05 40 SM] Til) N N ‘_ -1” "'45" Ec\ \ \ \ .1.4V 0 J. , r;/ l l l i l i 1] ii ;/// // -_ _-__a./// / —-— _...._L_ .— 3.0V 3.51! l l i 1 a 2”,, “.2.-.— l Lzuzuiaza' l i Ly “‘7‘; 7‘1‘6V ! / / z 4 /* [VT/’fizz-OSV / ///A /,/:// 3/ ///// Figure 10. Energy diagram of CuPc(SO'3'Na+)h oxidation in H20 (pH h) at SnO2 and Ti02. 41 through which electron transfer can occur. As the carrier concentration increases, the charge depletion region at the surface decreases in thickness. When the carrier density is sufficiently high enough, this depletion thickness will be 10 A0 or less, which enables electron tunneling to occur (5h). The redox couples of the dye and ferrocene in DMSO at plati- num are shown in Figure 9 for reference. The redox couples at SnO2 and Ti02 are shifted cathod- ically from the potential at which they occur at pltinum. When the number of charge carriers at the electrodes sur- face approaches Nc, degeneracy begins and electrode behavior becomes more metal like. Thus, the higher the carrier den- sity, the more metal-like a semiconductor electrode's be- havior is, i.e. reductions of CuPc(SOSNa+)u occur at cathodic potentials closer to potentials corresponding to the reduc- tions at platinum (8). The couples at SnO2 are shifted less and the carrier density higher indicating the SnO2 is more metal-like than the Ti02. Adsorbed Dye Studies SpectrOphotometric Studies of CuPc on SnO9 An absorption spectra of copper phthalocyanine (CuPc) deposited on SnO2 is shown in Figure 11. Absorption maxima occurs at 69hnm and 620nm in comparison to 691nm and 62hnm for CuPc deposited on glass (29). The spectra indicated the films to be in the at form. The CuPc thin films could be vacuum deposited on the SnO2 surface in one of three ways: 42 CuPc on SM] 2 I\ I: 3 O r t: a [1 c e Sn02 l l l T 400 550 750 Wavelength (n m) Figure 11. Absorption spectra of clean SnO2 and CuPc sub- limed on SnO2 electrodes. 43 with the crystallites parallel with the surface, with the crystallites standing obiquely to it, or some combination of these two arrangements (63). Assuming an electrode area of 0.8 cm2, a dye diameter of 1h.3 A0, and an extinction coef- ficient of 3.32 x iohm'1cm'1 at 618nm (27), one could de- termine the approximate thickness of the films from Beer's Law (A = be). The SnO contribution to the adsorption at 2 618nm was subtracted before calculating the thickness of the films. This method resulted in films varying between 20-100 monolayers with 3.91 x 1013 molecules per monolayer. Electrochemistry of CuPc on SnO2 The electrochemistry of CuPc sublimed on Sn02 in DMSO is shown in Figure 12. Cathodic scans showed the first re- duction wave (1) at -0.75 volts vs AgRE, shifited 0.13 volts cathodically from the peak potential of the solution compo- nent (Figure 5). Scanning back anodically resulted in an ox- idation wave (2) at 1.h volts vs AgRE. Both waves appeared chemically irreversible. Upon scanning to the second reduc- tion peak (3) potential, a large symmetrical desorption cur- rent peak was observed, coincident with the visible loss of CuPc from the electrode surface. Returning anodic scans in- dicated the reversible oxidation of the two reduction inter- mediates, (3',1') but a resorption process wasn't observed. If the potential was cycled repeatly between the cathodic and anodic limits, the desorption occurring at the second catho- dic peak potential slowly decayed away, the oxidation wave 44 38mv/sec i_i?200mv 2 3JuA/cm (a) 2' I— r 2 AI t—dSOOmv IRMA/cm2 3 38mv/sec (b) 3 l-——'l 50011” I I 8.3uA/cm2 75mv/sec (C) T 2 Figure 12. Cyclic voltammetry of CuPc sublimed on SnO2 in 0.1,; TEAP/DMSO. 45 (2) and two reduction waves (1,3) retained a constant magn- itude indicating the fact that some of the dye remained after the desorption process and was electrochemically active. All three peaks appeared to be chemically irreversible, i.e. dis- proportional peak current ratios. These electrochemical experiments indicate that the sub- limation process produces a tightly-bound form of dye and a loosely-held form which can be desorbed at a potential very near the second reduction. The formation of the dianion Spe- cies may force this desorption. If the dye remained purely surface bound at all potentials, electron transfer to and from it should occur with symmetrical reduction and oxidation peaks with a‘QSEpeak near zero. Oxidation of the reduction intermediates should be diffusion controlled since they were still near the electrode surface immediately after being de- sorbed. Diffusion control is indicated by the cyclic be- havior observed, i.e. [iEpeak23100mv. The cyclic voltammetric studies of CuPc sublimed on SnO2 in H20 resulted in poorly defined and irreversible reduction and oxidation processes. No desorption of the dye was ob- served during the scans in any aqueous media. ESCA of CuPc on SnO9 To facilitate the understanding of the adsorbed CuPc/ SnO2 electrochemistry, ESCA analysis of electrodes in various states of pretreatment along with standards was carried out. Changes in surface concentration and valence states of the 46 adsorbed dye occurring as a result of these electrochemical pretreatments are easily observed with ESCA. The Cu(2p1/2, 3/2) ESCA spectra are shown in Figure 13. Since Sn02 was a common component for all these materials, the binding ener- gies of all components were corrected to the Sn(3d5/2) line of standard Sn02, h86.2ev (h9). The Cu(2p1/2,3/2) peaks ex— cept for spectrum (g)(H20) are accompanied by multiplets at binding energies approximately 9-10 ev higher than the 2p transitions. This type of multiplet Splitting has been previously reported for metal (paramagnetic) oxides. This splitting has been contributed to the presence of oxygen in the oxide lattice (Sn02), adsorbed oxygen, or another oxygen containing molecule (55). Binding energies of the Cu(2p3/2) and Cu(Auger) peaks for the various samples are given in Table 1. Because there is little binding energy shigt in the Cu (2p1/2’3/2) peaks for the various copper species, the most useful information can be obtained from the Cu(2p3/2) - Cu (Auger) binding energies. This binding energy difference does shift with a change in the copper valence state. CuPc/ SnO2 pressed pellet, CuPc/SnO2 unused electrode, and CuPc/ SnO2 electrode (DMSO) [NBE'S (597.1ev, 596.8ev, and 597.0ev, respectively) correlate well with that of CuO (596.9ev) (57), thus indicating the presence of Cu(II) species. The KNEE for CuSOu (anhydrous)/Sn02 standard pellet is slightly lower, 595.8ev. The absence of multiplet splitting and the BE (59h.9ev) for the CuPc/SnO electrode in H20 seems to sig- 2 nify the presence of a copper species more reduced than Figure 13. 47 Cu(2p1/2 3/2) ESCA spectra of (a) CuSOu/Snoz powder pressed pellet, (b) CuPc/SnO2 powder pressed pellet, (c) CuPc/SnO2 unused electrode, (d) CuPc/SnO2 electrode in DMSO (scanned past 1st reduction peak and back anodically to the oxidation peak), (e) CuPc/SnO2 electrode in DMSO (scanned past 2nd reduction peak and back anodically to the oxidation peak), (f) CuPc/ SnO2 electrode in DMSO (scanned until desorp- tion process complete), (g) CuPc/SnO2 electrode in H20 (scanned until desorption process complete). Cu (2P1/2,3/2) N(E) 1300 cts o——-—4 500 cts 1 f) iWcm 9) WW 1300 C15 300 Cts fl/ (1) W 00 cts if 1300 cls l l 970 945 920 BINDING ENERGY (eV) 49 on. as. oz. ADV O z Anmv oogopomom Aomv monogamomA m.mqm o.mom 0.50m o.®om leessamazv.m.m on on on Ir on Av ea mofiwmocm mcwpcfim AsowsfiufiuCoo mmx opfipuaop mom popoomnoo a Loo Aodv oogopohom .>om.om: u Am\mpmvgm 0p popooaaooAav m.mom so m.oom omso o.oom one Acvm.mmo one omm\ooosp o.:om xe.oeovo.emm w.mmo -ooao mocm\omso omxo\oooae o.eom Ae.oeove.emm e.smo -ooao mosm\omso pandas coon» m.oom Am.oeovm.emm 0.3mm -ooHo mocm\omso popzoa Nocm\uoHHon e.eom Ae.oeovs.emm m.:mo cascades omso e.mem he.o.ees.emm e.mme soaso .m.mnw xovxxsoms 00. 54 vs Ag/AgCl reference. A maximum value of 1.5 }LA/cm2 was ob- tained at approximately 0.9 volts vs reference after which the photocurrent response decayed back to essentially zero. In another experiment, the light source was chopped at 13 Hz and the photocurrent measured using the lockin tech- nique. This technique employed the light chopper frequency as a reference for the lockin amplifier, thus allowing easier detection of the electrode photocurrent response. The photo- current as a function of anodic bias potential was studied. The results for two electrodes having approximately the same dye coverage are shown in Figure 16. Electrode A was biased at 0.2 volts, 0.5 volts, 0.9 volts, then 0.5 volts and 0.2 volts vs Ag/AgCl reference and the photocurrent response re- corded. The response appeared to reach a maximum between 0.5 volts and 0.9 volts. The second readouts at 0.5 volts and 0.2 volts were slightly less than the final steady response at 0.9 volts. The other electrode (B) was biased at 0.9 volts and an initial reading taken. The response decayed over a period of several minutes to a steady photocurrent approximately one- fourth of the initial value. As with electrode A, the photo- current response at 0.5 volts and 0.2 volts was slightly less than the final steady 0.9 volts response. After the photocurrent studies, visual examination of the electrodes indicated partial desorption and/or decomposi- tion of the CuPc film. This phenomenon appeared to be poten- tial dependent; it occured at potentials anodic of 0.5 volts. 55 Figure 16. Photocurrent response vs anodic bias potential plots of two CuPc/SnO electrodes in 0.05M Na C2Ou/DH 7 buffer. 2 2 56 IR filter UV filter LP47 filter (.1354 filter 93 monolayers Lockin sensitivity SOO hv .14. ( l J J 2.- i 2 12 PC. response . 0 ll (V) i i UWQ 9 I0 .8 .6 .4 .2 .0 Bios (v) 90 monolayers (9 4 43 ‘2 Dc. response 1 -«I (V) u— +0 0 (0 .8 .6 .4 .2 0 Bios (v) 57 As the bias potential was moved anodically, the photocurrent response increased. This would be indicative of a higher number of excited dye molecules and faster rate of electron flow from the dye to the semiconductor conduction band. As a result, the supersensitizing agent (oxalate ion) concentra- tion at the dye-solution interface could be rapidly depleted since diffusion of oxalate ion to the interface is not fast enough to maintain a high concentration. Thus, there is a build-up in the CuPc+ species concentration which with the bias potential and light probably enhance the desorption and/ or decomposition process. Cyclic voltammetry studies indi- cated a tightly-bound form of CuPc and a more loosely-bound form on the electrode surface. Since a steady photocurrent response was noted after the desorption—decomposition process, it seems the tightly-bound form of dye is retained on the surface. The photocurrent response then appeared to be less potential dependent. Increasing the oxalate ion concentration appeared to slow down the desorption-decomposition process. Stirring the solution could also retard this process by increasing trans- port of oxalate ion to the interface. However, the magnetic stirrer frequency coupled with the chopper frequency. This resulted in noise which obscurred the photocurrent response. The actual composition of the retained or tightly-bound layers on the electrodes is not known. Assuming the layers are essentially the same in the electrochemical and photo- chemical studies, ESCA studies appear to indicate the presence 58 of an electrochemically active species. It is possible the retained layers are not de-metalled phthalocyanine, but a decomposed form of CuPc; this decomposed form is electroch- emically active, i.e. contains double bonds. ggyalent Attachment Studies CuPc was attached to SnO2 via sulfonamide formation using gamma-aminopropyl-triethoxysilane or N,beta-aminoethyl- gamma-aminopropyl-trimethoxysilane and the sulfonyl chloride form of the dye (Figure 3). The sulfonamide procedure in- volves loss of HCl and the formation of a nitrogen-sulfur bond. Attachment via a thiol formation using mercapto- prOpyl-trimethoxysilane and tetraiodated CuPc (Figure 3) was unsuccessful because the dye was insoluble in aqueous media. The reaction should involve loss of HI and the formation of a carbon-sulfur bond between the dye and silane. Since CuPth was soluble in pyridine, either it or water/pyridine mixtures were also tried without success. The thiol form- ation has previously worked only in pH 7 buffer with water soluble dyes (58). Spectrophotometric Studies of Covalently-Attached CuPc An absorption spectrum of CuPc covalently-attached to SnO2 is shown in Figure 17. Absorption maxima were at 620nm and 69hnm corresponding to those of the dye sublimed on Sn02. Absorption due to the dye (0.02h) indicated approx- imately 2.h5 x 1013 dye molecules present on the surface 59 CuPc--Sn0 2 A b s o r b a n c e Sn02 _i l T l 400 550 750 Wavelength tnm) Figure 17. Absorption spectra of clean Sn0 and covalently- attached CuPc-SnO2 electrodes. 60 assuming a dye diameter of 1h.3 A0, extinction coefficient -1 (27), and a 0.5 cm2 sample area. If a 2 of 3.h x 10uM-1cm SnO2 molecule area was 30 A0 and the dye was arranged perpendicular to the surface, there should be, at most, eight dye molecules for every 550 A02 of surface, or approx- imately 7.35 x 1013 molecules/0.5 cm2. Approximately 1.03 x 1013 molecules/0.5 cm2 would be present if the dye mole- cules were arranged parallel with the surface. The actual dye arrangement is probably some combination of the above mentioned. The extent of silane coverage, number of active dye attachment sites, and steric hinderance limit the dye coverage. Electrochemistry of Covalently-Attached CuPc-SnO2 Electrodes The cyclic voltammetry of a chemically-modified elec- trode in DMSO is shown in Figure 18. Scanning cathodically (cyclic (b)), three reduction waves were abserved at -0.55 volts, -1.25 volts, and -1.80 volts vs AgRE. The peaks appeared chemically irreversible with three anodic waves at -0.58 volts, -0.98 volts, and -1.6 volts corresponding to oxidations of the reduction products. An irreversible oxidation of the dye occurred at 1.h volts vs AgRE. Stir- ring the electrolyte solution then scanning cathodically again, resulted in cyclic (c); two poorly resolved reduc- tions and an oxidation of the dye were observed. These waves maintained constant magnitude indicating the presence of an electrochemically active species. 61 Figure 18. Cyclic voltammetry of covalently-attached CuPc- SnO2 electrode in O'1flL TEAP/DMSO. 62 l-—-—-l 500mv I :Hul/cm2 38mV/sec (a) 7.4mV/sec (b) 38mv/sec (c) 63 Cyclic (c) of the covalently-attached dye is quite similar to cyclic (c) of the adsorbed dye (Figure 12). It is conceivable the chemically irreversible waves noted are really the catalytic oxidation and reduction of molecular oxygen since it is well known the phthalocyanines are good catalysts (59). The presence of 10'uM - 10'6M oxygen would be enough for the adsorbed or covalently-attached dye to catalyze its electrochemical reduction and oxidation. It is also possible the cyclic corresponds to the oxidation and reduction of a decomposed form of CuPc. The above spectrophotometric and electrochemical results do not totally confirm the success of the covalent attachment. Therefore, unsilanized SnOZ and TiO2 were used as controls to insure the results were not those of adsorbed dye. Unsilanized SnO2 was allowed to react with the sulf- onyl chloride form of the dye. It was then extracted in benzene and water in the manner the derivatized electrodes were. Spectrophotometric studies indicated a clean SnO2 surface; cyclic voltammetry of the electrodes also confirmed the absence of any adsorbed dye (Figure 18 (a)). It is conceivable that the silanized surface of the SnO2 could enhance adsorption. However, if the dye was adsorbed, extraction with water would decompose the sulfonyl chloride sites on the CuPc to form tetrasulfonated CuPc which is soluble. Therefore, the dye is most probably covalently-attached to the semiconductor electrode. 64 Photocurrent Response The photocurrent response vs anodic bias potential plot of a sulfonamide-linked CuPc-SnO electrode in 0.05M 2 sodium oxalate/pH 7 buffer is shown in Figure 19. The re- sponse appears to be potential dependent with its initia- tion beginning at approximately 0.5 volts vs Ag/AgCl ref- erence electrode. The photocurrent at any given anodic potential was steady over a period of several minutes, and is comparable to the photocurrent response of the adsorbed dye electrode after the desorption/decomposition process is complete. The covalently-attached dye electrode response was approximately one-fourth the final adsorbed dye electrode photocurrent response. Further conclusions concerning the stability and photocurrent response of the covalently-at- tached dye can not be made without more experimentation. 65 Figure 19. Photocurrent response vs anodic bias potential of a covalently-attached CuPc-SnO electrode in 0.05M Na2020u/pH 7 buffer. 2 NT «I. 1.0 I 0“” 66 LOCKIN SENSITIVITY lfllluV IR FILTER UV FILTER LP 47 FILTER LP 54 FILTER J_ j .6 ,4 Bias (volt: s) 1)- 0 ID" P.c. (volt 3) b6 . CHAPTER IV SUGGESTIONS FOR FUTURE WORK 68 Future work would consist of several closely related investigations. Better methods of covalent-attachment to increase the phthalocyanine surface coverage would be studied. Efficiency of the photosensitizing process could be increased by using electrodes with several monolayers of covalently-attached dye. This could be accomplished by two methods: (1) coupling of phthalocyanine molecules with cyanuric chloride (60), or (2) coupling of a silicon phthalocyanine derivative with itself (61,62). Extensive photocurrent studies of the adsorbed and covalently-attached dye electrodes would be carried out. Use of a pulse dye laser would increase the light inten- sity and photocurrents. Photocurrent response studies using different supersensitizing agents, light aging studies with the chemically-modified electrodes biased at different potentials, and ESCA studies of electrodes in various states of electrochemical and photocurrent treatment would aid in a better understanding of the chemical stability of the dye and kinetics of the photo- sensitizing process. "LI ST OF REFERENCES" 10. 11. 12. 130 1h. 70 "LIST OF REFERENCES" M.S. Wrighton, D.S. Gimley, P.T. Wolczonski, A.B. Ellis, D.L. Morse, and A. Linz, Proc. Nat'l Acd. Sci., U.S.A., 1a, 1518 (1975). H. Gerischer, Photochemistry and Photobiology, 16, 2h3 (1977). __ S. Mishitsuka and K. Tamsru, J. Chem. Soc., London, 13, 705 (1977). - T. Osa and M. Fujihira, Nature, 26h, 3R9 (1976). D. Hawn and N.R. Armstrong, unpublished results (man- uscript in preparation). T. Osa and T. Kuwana, J. Electroanal. Chem., 22, 389 (1969). - V.A. Myamlin and Y.V. Pleskov, "Electrochemistry of Semiconductors," Plenum Press, New York, 1967. S. N. Frank and A.J. Band, J. Amer. Chem. Soc., _1, 7h27 (1975)- K.G. Allum, R. Hancock, and P. Robinson, J. Opganomet. G.B. Harper, Anal. Chem., g1, 3h8 (1975). C.M. Elliott and R.W. Murray, Anal. Chem.,gg, 12h? (1976). N.R. Armstrong, A.W.C. Lin, M. Fujihara, and T. Kuwana, Anal. Chem., pg, 7h1 (1976). P. R. Moser, L. Wier, and R. W. Murray, Anal. Chem., Q_, 1882 (1975) P.R. Moser and R.W. Murray, J. Amer. Chem. Soc., in press. 71 15. B.F. Watkins, J.R. Behling, E. Koriv, and L.L. Miller, J. Amer. Chem. Soc., 31, 3su9 (1975). 16. D.G. Davis and R.W. Murray, Anal. Chem., 82, 19h (1977). 17. K. Siegbahn, C. Vordling and A. Fahlmann, et al., "Elec- tron Spectroscopy for Chemical Analysis," Almquist and Wilkells, Uppsala, Sweden, 1967. 18. K. Siegbahn, C. Vordling, K. Hamru, and J. Redman, et al., "ESCA-Atomic, Molecular and Solid State Structures Studied by Means of Electron Spectroscopy," Almquist and Wilksells, Uppsada, Sweden, 1967. 19. Kane and Larrabee, "Characterization of Solid Surfaces," Plenum Press, New York, 197R. 20. J.T. Yates, Chem. and Eng. News, Aug. 26, (197k). 21. J.T. Yates and T.E. Madey, Surface Sci., 8;, 257 (197A). 22. K.S. Kim, T.J. O'Leary, and N. Winograd, Anal. Chem., 23. M.J. Dressers and T.E. Madey, Surface Sci., 88. 533 (197h). 2h. A.W.C. Lin, N.R. Armstrong, and T. Kuwana, Anal. Chem., in press. 25. J.M. Bolts and M.S. Wrighton, J. Phys. Chem., 89, 26h1 (1976 . - 26. R. Memming, Photochem. and Photobiology, 18, 325 (1972). 27. F.H. Moser and A.L. Thomas, "Phthalocyanine Compounds," ACS Monograph Series, No. 157 (1963). 28. D.H. Busch, J.R. Weber, D.H. Williams and N.J. Ross, J. Amer. Chem. Soc., 88, 5161 (196k). 29. E.?. ggpia and F.D. Verderame, J. Chem. Phys., 8, 267R 19 . “ 30. L.D. Rollman and R.T. Iwamot, J. Amer. Chem. Soc., 28, 31. A.N. Sidorov, Russian J. Structural Chem., 18, 255 (1973). 33. 3h- 35. 36. 37. 38. 39. h0. M1. 1.2. 43. AS. h6. h7. R8. 72 E.A. Rodyushkuna, B.D. Bereyin, and N.R. Tarasevich, Elektrokhimiya, 2, R10 (1973). Y.M. Stolovitskii, A.Y. Shkuropatov, and V.V. Shichhov, Biofizika, 12, 820 (197A). V.S. Shumov and M. Heyrovsky, J. Electranal. Chem., 88, h69 (1975). -— G.A. Alferov and V.I. Sevast'yanov, Elektrokhimiya, 827 (1975). J.P. Randin, Electrochimica Acta., 19, 83 (197h). J D 11 T.A. Maas, M. Kuifier, and J. Zwart, Comm., 88 (1976). . Chem. Soc., Chem. Y. Niwa, H. Kobayaashi, and T. Tsuchiya, J. Chem. Phys., M. Zeller and R.G. Hayes, J. Amer. Chem. Soc., 25, 3855 (1973)- ‘- J.H. Weber and D.H. Busch, Inorg. Chem., 8, h69 (1965). R.W. Higgins and C.L. Hilton, J. Org. Chem., 18, 1275 (1951). - Japan Patent No. 19,225. W.F. Kosoncky, S.E. Harrision and R. Stander, J. Chem. M. Babai, N. Tshermikooski, and E. Gileadi, J. Electro- chem. Soc., 119, 1018 (1972). R.K. Quinn and N.R. Armstrong, J. Vac. Sci. Technol., 18, D. Sawyer and J. Roberts, "Experimental Electrochemistry For Chemists," John Wiliz and Sons, New York, 197k. Y.S. Shumov, V.I. Sevast'yanov, and 0.0. Komissarov, Biofizika, 18, RB (1973). A. Vogel, "Practical Organic Chemistry," 2nd ed., Wiley, New York, 1966. R9. 50. 51. 52. 53. 5h. 55. 56. S7. 58. S9. 60. 61. 62. 63. 73 P.A. Grutsch, M.V. Zeller, and T.P. Fehlner, Inorg. Chem., 2’ 11431 (1973). N.R. Armstrong and T.V. Atkison, private communication. K. Bernauer and S. Fallab, Helvetica Chemica Acta, 88, 1287 (1962). D.W. Clac, J.B. Raynor, L.P. Stodulski, and N.R. Symons, J. Chem. Soc. A, 997, (1969). W.P. Gomes and F. Cardon, Z. ngs. Chem., 88, 333 (1973). N.R. Armstrong, private communication. T. Novakov and R. Prins, in "Electron Spectroscopy," Proc. Intern., Conf. Held at Asilomar, Calif. 7-10 Sept. 1971, Ed. D.A. Shirley, Worth-Holland, Amster- dam, 1972. P.A. Larson, J. Elec. Spect. and Related Phenom., 8, 213 (197h)o G. Schon, Surface Science, 35, 96 (1973). 8.8. Hasenoff, Cand. J. Biochem., 82, 7h2 (1971). J. .R. Go§denstein and A.C. C. Tseung, Nature, 222, 869 1971 -—— K. Venkatasaman, " The Chemistry of Synthetic Dyes," Academic Press, New York, 1972. Japan Patent 5,711. J.N. Esposito, J. E. Lloyd, and M.E. Kenney, Inorg. Chem., 5. 1979 (1966). R.S. Kirk, Molecular Crystals, 5, 211 (1968). l Ill-II | Iii Ii! i III I