USE OF THE DIFFERENTIAL REFLECTOMETER IN THE STUDY OF THIN FILM CORROSION PRODUCTS ON COPPER By CHARLES WILLIAM SHANLEY A DISSERTATION PRESENTED TO THE GRADUATE COUNCIL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS OF THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 1977 ACKNOWLEDGEMENTS The author wishes to express his sincere appreciation to his advisor, Dr. Rolf E. Hummel, for his guidance and confidence throughout this work. Special thanks are extended to Dr. J. B. Andrews for his help in the design and construction of much of the apparatus used in this work, and to Dr. R. J. Nastasi -Andrews for numerous stimulating discussions. Special thanks are also extended to Dr. E. D. Verink, Jr., for his advice concerning the electrochemical aspects of this research, and to Drs. J. J. Hren, R. Pepinsky, D. B. Dove, and E. D. Verink, Jr., for serving on the author's supervisory- committee . The author wishes to express his appreciation to Dr. B. Cavin, who performed the X-ray analyses mentioned in this work, and to Mr. Patrick Russell, who performed the ellipsometric measurement. The National Science Foundation provided financial support for this research, which is gratefully acknowledged. TABLE OF CONTENTS Page ACKNOWLEDGEMENTS i i LIST OF TABLES v LIST OF FIGURES vi ABSTRACT x SECTION I INTRODUCTION 1 II LITERATURE REVIEW 11 Introduction 11 Related Techniques 12 Cuprous Oxide 28 Band Structure Calculations 28 Optical Spectroscopy 31 Growth and Kinetics 37 Summary 41 III EXPERIMENTAL PROCEDURE 46 Introduction 46 Sample Preparation 47 Metallographic Preparation 47 Special Procedures for In Situ Samples. 50 Electrolyte Solutions 54 The Differential Reflectometer 54 Introduction 54 The Instrument 56 Advantages and Disadvantages 59 Other Equipment 62 Operating Conditions 66 Other Considerations 67 Summary 70 TABLE OF CONTENTS - continued. SECTION Page IV RESULTS AND DISCUSSION 72 Introduction 72 Theory of Differential Ref lectograms of Thin Films 74 Non-7n Situ Experiments 95 In Situ Experiments 10 7 Further Discussion 116 V CONCLUSIONS AND SUGGESTIONS FOR FURTHER WORK 124 REFERENCES 127 BIOGRAPHICAL SKETCH 131 LIST OF TABLES Table Page 1 Comparison of Peak Positions for Absorption and Reflection Spectra 32 2 Major Peak Positions in C112O Optical Spectra Compared to Transitions Calculated from the Band Diagram 45 3 Composition of Electrolyte Used in Nil Chloride Experiments 55 4 Optical Constants ft = n+ik for Copper and Cupric Oxide (CuO) as a Function of Wavelength 83 5 Major Structure in CU2O Spectra Obtained from Other Techniques Compared to Differential Ref lectometry 120 LIST OF FIGURES Figure Page 1 Typical experimental arrangement for ellipse-metric measurements. The angle <{> is the angle of incidence 13 2 Schematic diagram of a modulated ellipsometer developed by Jasperson et al. 14 The detector (DET) sends the modulated signal to two lock- in amplifiers tuned to 100 KHz and 50 KHz and to a low pass filter. The data is reduced by the computer, C, and displayed on a digital meter, DV. Q' or A' is an additional polarizer used to calibrate the instrument... 17 3 Electrochemical modulation apparatus used by Mclntyre.22. The HIS 60 70 is a computer used to reduce the data 22 4 Schematic diagram of a rapid scanning spectrophotometer used by Mclntyre et al . See text for details 25 5 Band structure diagram for cuprous oxide (Cu20) calculated by Dahl and Switendick 30 6 Reflectivity spectrum of polished and polished and etched Cu?0 crystals 33 7 Electroref lectance spectrum of single crystals of CU2O for various AC voltages superimposed on a 1 KV DC bias 36 8 Electroref lectance spectrum of CU2O at 85 K and no DC bias. The modulating electric field was 60 KV/cm 38 9 Schematic diagram of the differential ref lectometer 57 10 Schematic diagram of the standard corrosion cell used for non-in situ experiments 63 LIST OF FIGURES - continued. Figure 11 12 Schematic diagram of the corrosion cell used for in situ experiments. Details of the sample holder are not shown for clarity Theoretical (Figure 12A) and experimental (Figure 12B) Pourbaix diagrams for copper in nil chloride solutions 65 73 15 14 15 Differential reflectogram of a sample prepared in the CU2O region of the Pourbaix diagram (-225 mv SCE, pH 9.2, 66 hours) Differential reflectogram of a sample prepared in the CuO region of the Pourbaix diagram (+200 mv SCE, pH 9.2, 18 hours) Typical compositional modulation differential reflectogram. One sample is pure copper, the other copper with 1 at . % aluminum in solid solution 75 76 77 16 17 19 Schematic representation of a typical sample with an oxide film covering half the exposed surface 79 Variation in AR/R as a function of film thickness for a CuO film on a copper substrate. The wavelength of light used is assumed to be 600 nm 82 Theoretically constructed differential reflectogram of a CuO film 100 A thick on a copper substrate. A = 600 nm 85 Theoretically constructed differential reflectogram of a CuO film 35 A thick on a copper substrate. X = 600 nm 87 LIST OF FIGURES - continued Figure Page 20 Theoretically constructed differential reflectogram of a CuO film 500 A thick on a copper substrate. A = 600 nm 88 21 Theoretically constructed differential reflectogram of a CuO film 7500 A thick on a copper substrate. X = 600 nm. 22 Variation in AR/R as a function of film thickness for a CuO film on a copper o substrate over the range d = 0 to 200 A. The wavelength of light is assumed to be 600 nm 91 23 Comparison of the differential reflectogram of an experimentally prepared CuO film (potentiostated at +500 mv SCE at pH 9.2 for 18 hours) to a theoretically constructed reflectogram of a CuO film 25 A thick 93 24 Differential reflectogram of a sample prepared by immersion in distilled water for eighteen hours, with a continuous stream of oxygen bubbled through the water. The structure is characteristic of a mixed oxide of CU2O and CuO 100 25 Differential ref lectograms of three areas of the same sample, which had been potentiostated near the CU2O/C11O equilibrium line (-50 mv SCE at pH 9.2) for eighteen hours 102 26 Differential reflectogram of a sample potentiostated in the CuO region (A) of the Pourbaix diagram, and of the same sample after the potential has been changed to the CU2O region (B) . Curve C is the same sample potentiostated for a longer period of time in the CU2O region. The baselines of the curves have been shifted for clarity 105 LIST OF FIGURES - continued. Figure Page 27 Differential ref lectograms of two samples potentios tated under the same conditions (5.5 hours at -200 mv SCE, pH 9.2). The first was run in the non-in situ cell and measured in air, while the second was prepared and run in situ 108 28 Differential ref lectograms of a sample half covered with a Cu20 film, showing the sample as measured in air (A) , in the empty in situ cell (B) , and in the in situ cell filled with electrolyte (C) Ill 29 In situ differential ref lectograms of a copper sample potentiostated at -200 mv SCE, pH 9.2, which is in the Cu20 region of the Pourbaix diagram 113 30 Continuation of the in situ experiment of Figure 29. Note the change in scale. Since the curves do not change appreciably during these times, the zero lines have been shifted for clarity 114 Abstract of Dissertation Presented to the Graduate Council of the University of Florida in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy USE OF THE DIFFERENTIAL REFLECTOMETER IN THE STUDY OF THIN FILM CORROSION PRODUCTS ON COPPER By Charles William Shanley December 1977 Chairman: R. E. Hummel Major Department: Materials Science and Engineering An investigation was undertaken to examine the feasibility of using the technique of differential ref lectometry to identify and study thin film corrosion products on metal substrates. Differential ref lectometry is a form of modulation spectroscopy which measures the normalized difference in reflectivity between two samples, or two adjacent areas of the same sample. In these experiments, one half of each sample was protected by an insulating lacquer while the other half was allowed to form a thin film corrosion product. When the lacquer was subsequently removed, the normalized difference in reflectivity as a function of wavelength was measured over the range A = 800 nm to 200 nm (1.55 eV to 6.20 eV) . The use of a transparent lacquer permits monitoring of the film in situ. Because of the differential nature of the technique, the effect of the metal substrate is subtracted out, so that the resulting differential reflectogram is related to the absorption spectrum of the corrosion product. The similarity in the structure between reflection, absorption, and differential reflectivity spectra substantiates this viewpoint. In these experiments, copper was chosen as the metal on which the thin film corrosion products were grown. Characteristic spectra for the copper oxides Cu~0 and CuO were obtained, and were used to help define the boundaries of the regions of stability of the copper-water Pourbaix diagram. In addition, the evolution with time of a Cu?0 film was monitored in situ. It was established that in the thickness range of o passive and semipassive films (d = 0 to 200 A) dif- ferential ref lectometry produces a unique spectrum for each corrosion product, and that the intensity of the structure in the differential reflectogram is related to the thickness of the film. A method is described for calculating the differential ref lectograms of thin film corrosion products for various film thicknesses if the optical constants of the film and substrate are known. Sample spectra are calculated for various thicknesses of a CuO film on a copper substrate. These spectra compare favorably with experimental curves. XII SECTION I INTRODUCTION Although metallic corrosion is a familiar phenomenon, many aspects of the process remain poorly understood. The reason is not that there is some fundamental bit of in- formation which is lacking; rather, it is that corrosion itself is not a single process. Generally speaking, there are eight basic forms of corrosion, any one or more of which may be present at the same time. The effects may be additive, as when pitting and uniform corrosion attack a holding tank, or even synergistic, as when erosion corro- sion and selective leaching attack a brass pipe. In addi- tion, matters taken to eliminate one form of corrosion may inadvertantly introduce new problems, as when the painting of the inside of a tank with a protective coating to eliminate uniform corrosion induces pitting in spots where the coating is imperfect. Altogether, the cost of corro- sion in the United States alone is estimated to be in 2 excess of eighty billion dollars annually. In order to correct a corrosion problem, an engineer must know what kind of corrosion process is present. Fre- quently, the answer to this question is based on the identification of the corrosion product. For bulk speci- mens with massive corrosion, x-ray diffraction techniques, perhaps supplemented by wet chemical analysis, are ordi- narily sufficient. Many such bulk corrosion problems are long familiar to the corrosion engineer and may be identi- fied by inspection. For thin films the identification of the corrosion products is considerably more difficult. It is not uncommon for thin, semipassive layers of corrosion products to form on metals such as aluminum, copper, or silver. Such layers, whether they be oxides, hydroxides, chlorides, or something more complex, are in general insulating or at best semiconducting. Since the aforementioned metals are extensively used by the elec- tronics industry, these thin corrosion films, which form insulating or rectifying contacts on electrical equipment, are of great concern. As in the case of bulk corrosion, the identification of the corrosion product is the first step in solving the corrosion problem. Unfortunately, such films are frequently difficult to analyze. X-ray diffraction techniques, which probe as much as a millimeter into the surface of a specimen, will often not show up an oxide layer on the outer one hundred thousandth millimeter of the surface. A more surface sen- sitive, less penetrating form of spectroscopy is called for. Electron spectroscopy techniques seem to provide the answer. Auger Electron Spectroscopy (AES) and Electron Spectroscopy for Chemical Analysis (ESCA) are the most commonly used techniques. Both have certain advantages over x-ray techniques for the study of thin films. Auger analysis and ESCA penetrate only a few atomic distances o (typically around 25 A) into the sample. This makes them well-suited for thin film analysis. Auger spectroscopy is not particularly well-suited for the identification of specific compounds, but ESCA, as its name implies, can identify compounds as well as elements. If the AES or ESCA apparatus is equipped with an ion milling device, the evolution with depth of the thin film corrosion product may be also studied. Both techniques are precise, well- known, and the apparatus is readily available. No technique is without its limitations, and AES and ESCA are not exceptions to this rule. The equipment for either is expensive and complex, and requires trained operators. They are both destructive techniques, in that an electron or ion beam abrades away the sample. They are not particularly fast. However, their most fundamental problem is that they require the sample to be placed in a vacuum. Removing a sample from its normal environment for study is a common research technique, but it is never done without a certain amount of misgivings. The boundary conditions are changed, and the results may not be the same as if the study were performed in situ. If a ti- tanium or aluminum specimen with an oxide coating is placed into a vacuum chamber, there will probably be no change in the oxide layer. However, if a sample with a fragile hydroxide layer or a complex compound containing water of hydration is removed from an electrolyte and exposed to air for a few hours, it is likely that the compound will not remain the same. If the sample is - 7 placed in a vacuum chamber and evacuated to 10 torr, the probability of changes in the sample approaches certainty. Some analytical technique which does not require a vacuum, and can hopefully be used for in situ analysis, is clearly needed. Optical techniques may provide a solution in such cases. Many of these thin film corrosion products are semitransparent to the eye. This implies that the films preferentially absorb certain wavelengths of light in the visible region of the spectrum. This absorption could be the basis for the analysis of the thin films. In addi- tion, optical techniques have other advantages over x-ray or electron spectroscopy. A light beam penetrates typically fifty atomic dis- ci tances (around 100 A) into a metal. It penetrates a much further distance through a semitransparent film, making it suitable for examining the entire depth of the film with- out penetrating extensively into the metal substrate. In addition, optical techniques are non-destructive tech- niques. But the greatest advantage of optical techniques is that they do not require a vacuum; indeed, the proper technique can study a sample in situ. Traditionally, ellipsometry has been the optical technique used to study thin films on substrates. This technique measures the degree to which a film changes plane polarized light into elliptically polarized light, and the phase difference between the two components of the elliptically polarized light. Using this information, ellipsometry can calculate any two of the three para- meters n, k and d, where n is the index of refraction, k is the index of absorption, and d is the thickness of the film. Computerized iteration techniques permit the determination of all three parameters, but the process is not rapid. High speed computer -controlled ellipsometers , although expensive, have improved this situation some- what . 4 One technique, differential ref lectometry , provides fundamentally different information when compared to ellipsometry, or for that matter, to AES or ESCA. By monitoring the difference in reflectivity between two samples as a function of wavelength, a series of peaks is obtained, each of which represents a transition from one energy band to another in the sample. If one sample is clean metal and the other sample metal with a thin film of corrosion product on it, measurement of the difference in reflectivity should show only the effect of the corrosion product. Since each compound is presumed to have a unique set of energy levels, this technique can identify the presence of a corrosion product, its identity, and its evolution with time. The measurements may be performed in situ, if need be. In addition, it is nondestructive, fast, and does not require a vacuum. In addition to studies on existing corrosion prob- lems, the technique of differential ref lectometry is use- ful to corrosion engineers in other ways. Pourbaix dia- grams are often used to represent a large amount of thermodynamic data concerning corrosion in an easy-to- interpret form. Such a diagram is a plot of potential versus pH, and shows regions of thermodynamic stability for metals and their associated corrosion products. A Pourbaix diagram may be constructed from thermodynamic data on the metal and corrosion products with the aid of the Nernst equation, 0 nF Ti^dJ where E is the half-cell potential, En is the standard half-cell potential, R is the gas constant, T is the absolute temperature, n is the number of electrons transferred, F is the Faraday constant, and [a 1 and ' ' ' L ox [a j] are the activities (concentrations) of the oxidized l fed and reduced species. This equation is used to define regions of thermo- dynamic stability in the Pourbaix diagram. It is well- known, however, that many real physical processes do not take place under equilibrium conditions. Hence the equi- librium diagram may not apply. Therefore it is of great interest to the corrosion engineer as well as the thermo- dynamicist to construct experimental Pourbaix diagrams to which the thermodynamically correct Pourbaix diagram may be compared. This experimental diagram is a practical guide which may be useful in engineering applications as well as being a comparative measure of how closely experi- mental conditions approach equilibrium conditions. Experimental Pourbaix diagrams are normally con- structed using the potentiodynamic method. At a fixed value of pH, the potential of the sample relative to some convenient reference is slowly varied while the current through the sample (which is directly proportional to the corrosion rate) is monitored. A plot of applied potential versus the logarithm of the corrosion current is normally obtained on an x-y recorder. If the potentiodynamic scan is begun with the potential applied to the sample well into the immunity region of the Pourbaix diagram, the corrosion current should be essentially zero. As the potential becomes more cathodic and the sample moves out of the immunity region, the corrosion current should in- crease. Any abrupt change in the corrosion current sig- nals the formation or dissolution of a corrosion product. For example, a sudden decrease in the corrosion current of several orders of magnitude as the potential is made more cathodic may indicate the onset of passivity at that po- tential and pH. Numerous scans at different values of pH serve to establish the experimental Pourbaix diagram. One difficulty with this technique is that although the formation of a corrosion layer may be observed, the identity of that corrosion product is not determined. If the corrosion product forms in or near a region of the Pourbaix diagram where a certain product is expected it is not unreasonable to assign that product to that area of the experimental Pourbaix diagram. However, borderline cases or conflicting results are not uncommon, as is a good deal of experimental scatter. Independent verifica- tion of the corrosion product is essential. Specimens prepared potentiostatically are inherently closer to equilibrium than potent iodynamically varied samples, and may be analyzed to determine the identity of the corrosion product. Since many of these films are too thin for x-ray analysis, optical methods represent the only simple techniques for examining these films without resorting to a vacuum. Because differential ref lectometry seems to produce a unique set of peaks for each corrosion product, it should be possible to identify corrosion products pre- sent in various regions of the experimental Pourbaix diagrams. In addition, in situ techniques make possible the study of kinetics of film dissolution or formation. Initial studies of any new technique must be made on carefully selected samples. Copper was chosen as the metal upon which the films would be formed, since copper has fairly well- characterized corrosion products, most of which have a visible color which implies absorption bands in the visible region. In addition, theoretical and experimental Pourbaix diagrams have been established. It is the purpose of this work to investigate the feasibility of using the technique of differential reflec- tometry to identify and characterize thin film corrosion 10 products on metal substrates. In addition, the concept of using potentiostatic electrochemical techniques in con- junction with differential ref lectometry to delineate areas of stability in experimental Pourbaix diagrams is to be investigated. To this end, the advantages of dif- ferential ref lectometry , such as speed, freedom from vacuum requirements, and, most importantly, the possi- bility of true -in situ measurements , are used to supplement conventional potentiostatic and potentiodynamic research techniques. Differential ref lectometry provides a techni- que for the direct observation of corrosion phenomena, thereby supplementing the indirect technique of potentio- dynamic scans. In this respect it is similar to x-ray diffraction, which can provide a direct way of studying phase changes in materials which are indirectly indicated by res is tometric or differential thermal analysis (DTA) techniques . No single technique, or combination of techniques, can be applied to the study of all corrosion phenomena. It is believed, however, that differential ref lectometry , by supplementing the information obtained using more con- ventional techniques, will provide a rapid, reliable, and more complete means of studying thin film corrosion phenomena. SECTION II LITERATURE REVIEW Introduction In the development of a new analytical technique, two problems must be considered— the mechanics of the technique itself and the interpretation of the results of that technique. Therefore, this review will be divided into two major parts. The first is a discussion of several optical methods which have been developed for the study of thin films. Their chief advantages and disadvantages, as well as how they relate to differential ref lectometry , will be discussed. Most of these techniques are useful primarily for the study of very thin films and seek to find the optical constants n and k and the thickness, d, of the film. Differential ref lectometry seeks only the identity of the film and its thickness, without regard for the particular optical constants. The second part of this review will concentrate on optical studies of cuprous oxide, Cu^O. In order to identify thin film corrosion products with the dif- ferential ref lectometer , one must know the absorption spectra or optical constants of each corrosion product as 11 12 a function of wavelength. This initial study has con- centrated on the oxides of copper, Cu?0 and CuO. Of the two, C^O has been more extensively studied since it possesses an intrinsically interesting electronic struc- ture. Cuprous oxide, Cu20, has a rich exitonic spectrum as well as a fundamental energy gap in the visible area of the spectrum. (An exiton is a bound electron-hole pair produced by some semiconductors by the absorption of an appropriate wavelength of light.) Cupric oxide, CuO, has not been as extensively studied. Fortunately, one study which has been done provides the optical constants as a function of wavelength, something which has not been done for Cu?0. Related Techniques The traditional method for investigating thin films on substrates has been ellipsometry . In a typical experi- mental arrangement, shown in Figure 1, a monochromat ic light beam passes through a polarizer to produce plane polarized light. This light is reflected from the metal or metal/film surface at the angle of incidence, $ , pro- ducing elliptically polarized light. This elliptically polarized light may be visualized as two perpendicular plane polarized waves of amplitude ratio 4> and phase dif- ference A. These ellipsometric parameters are determined 13 bfl C o H c CD C C ■H 0) e m-i ■ H O M are then used to calculate the optical constants n and k. The situation is not quite as simple as this, of course. The values of A and \p by themselves are insuf- ficient to uniquely determine n, k, and d for the film. Until recently, the film thickness, d, had to be deter- mined independently, by coulometry, interf erometry , or a series of ellipsometric measurements. None of these methods are particularly easy, and all are time-con - 7 8 suming . However, Paik and Bockris have shown that the ratio of the intensity of the light reflected from the film parallel to the plane of incidence (ri i) to that polarized perpendicular to the plane of incidence (ri) can be used in addition to ijj and A to determine the unknowns n, k and d. Unfortunately, only a few existing automatic ellipsometers , such as those of Paik and Bockris and Cahan and Spanier, can easily measure these quantities. 15 9 in A new approach by O'Handley and O'Handley and Burge calculates the effective reflectivity from A and ip and uses this quantity in addition to A and v/j to find the three unknowns, n, k, and d, for the film. As O'Handley points out, two possible values of the film thickness, d, can be produced using this analysis, but since they often differ by an order of magnitude, only a general idea of the thickness of the film is necessary to unambiguously determine all three parameters. Another feature of the classic ellipsometer which has made it less than ideal for thin film investigations is its lack of speed. In electrochemical investigations, many reactions occur very fast, and the determination of A and ty for a classic ellipsometer may require many minutes. Modifications to the normal operation of an ellipsometer, such as off -null techniques, can increase the sensitivity to transients and facilitate the moni- toring of known films, but more exotic techniques are required for unknown films of very rapid film formation. One approach to the problem is to speed up classical ellipsometry . Such an approach has led to a series of automatic ellipsometers , such as that of Cahan and 3 \2 Spanier, and recently to a commercial instrument. These instruments, which often use a continuously rotating 16 analyzer and a dedicated minicomputer, can analyze a film 1 2 for n, k, and d in four to ten seconds. Their chief drawback is their expense, since a commercial automatic ellipsometer costs approximately $20,000, or about four 1 3 times the cost of a conventional ellipsometer. Various kinds of modulation techniques have also been employed to speed up ellipsometry . The modulated para- meter is ordinarily the polarization of the light beam, which is varied by a photoelastic crystal or by a Faraday cell. Alternately, an oscillating analyzer or polarizer may be used. Figure 2 shows one such modu- lated ellipsometer developed by Jasperson et al . It is similar to a standard ellipsometer except that instead of a normal compensator it uses a piezoelectrically -driven quartz crystal to modulate the phase shift of the light beam. The detector uses two lock-in amplifiers to simultaneously monitor the reflectivity at the modulation frequency (50 KHz) and at twice the modulation frequency (100 KHz). In addition, the DC reflectivity is also monitored using a low pass filter in order to take into account changes in source intensity, detector response, and so on. The 50 KHz and 100 KHz signals contain in- formation from which n and k for the film may be derived if the. thickness , d, is known. If it is not, the method 17 -SH Figure 2. Schematic diagram of a modulated ellipsometer developed by Jasperson e t a 1 . 1 4 The detector (DET) sends the modulated signal to two lock-in amplifiers tuned to 100 KHz and 50 KHz and to a low pass filter. The data is reduced by the computer, C, and displayed on a digital meter, DV. Q' or A' is an additional polarizer used to calibrate the instrument. 9 of O'Handley may be used to find n, k, and d simul- taneously. The instrument is said to be an order of magnitude more precise than standard ellipsometers and can measure A and >p in less than thirty seconds. Differential ref lectometry , as used in the study of thin film corrosion products, is similar to the related technique of differential reflection spectroscopy used by 17-19 Mclntyre and others. Differential reflection spec- troscopy, however, uses polarized light and seeks to find the optical constants ft = n + i k for the film, while differential ref lectometry does not require polarized light and seeks only the reflection spectrum of the film to identify it. As in differential ref lectometry , the differential reflection spectroscopy technique of Kolb and Mclntyre measures the normalized difference in reflectivity, AR = R1'R2 R R, where R-^ is the reflectivity of the metal and R? the re- flectivity of the metal/film composite. The difference in reflectivity is normalized by dividing by the re- flectivity of the metal rather than the average reflec- tivity. Additional information may be obtained by polarizing the light beam parallel to [(y),,] and per- 19 AD pendicular to [ (—} , ] , the plane of incidence. Mclntyre 18 and Aspnes have shown that if the film thickness is very small relative to the wavelength of light being used, then and rAR (TTJJ_ !tt d n„ costf) t, - e7 r ImC-Ur^) (1) (AR 1 R J 0 Itt d n0 cos* /V^x-^Y^CV^sin2^^ Imi- — 5—! 1 =s — ]}(2) A G0^2 l-(^_)(en+g9)sin" •0 2- where t)> is the angle of incidence. The surrounding phase (either air or electrolyte) has a real dielectric con- stant, en, and index of refraction, nn , while the thin absorbing film and substrate metal have complex dielectric constants, e, and &~, respectively. Using these equations, there are two methods for measuring the optical constants of the film. 1. Measure the two quantities {—) , and (y) , , at a single angle of incidence. 2. angles of incidence. The details of each kind of measurement have been given by Kolb and Mclntyre. Method 2 is not used be- cause of mathematical difficulties in extracting the Measure either (^-) , or (^) , , at two or more R 20 optical constants n and k. Method 1 has been used suc- ° 19 cessfully to study adsorbed layers on gold 6 A thick. This technique has certain advantages over conven- tional ellipsometry which are similar to the advantages of differential ref lectometry , such as easily performed measurements in situ, fast response time, and measure- ments over a range of wavelengths. However, it is a dif- ference technique rather than a differential technique because the reflectivity of the metal and the metal/film composite is not measured at the same time. The reflec- tivity of the clean metal surface is measured in air, rather than in situ. The sample cell is filled with electrolyte and the sample potentiostated in the desired region of the Pourbaix diagram. After an appropriate amount of time, the reflectivity is again measured as a function of wavelength for both polarizations of light. The normalized difference in reflectivity is then calcu- lated point by point by computer. Several points need to be considered when comparing differential reflectance spectroscopy to differential ref lectometry . In differential reflectance spectroscopy, two scans over the entire wavelength range are necessary to determine the complex dielectric coefficient of the film, in addition to the reflectivity measurements and 21 Kramers -Kronig calculation necessary to obtain the optical 1 o constants of the substrate. ' Secondly, the output is not a continuous curve, but only a series of points. Sharp changes can be missed. Thirdly, if the film is absorbing, the thickness must be determined independently before n and k for the film can be obtained. Finally, I o equations 1 and 2 are only accurate if d/X < .003. For A = 250 nm (approximately the lowest wavelength used in o this work) d must be less than 7.5 A. Hence the technique is useful mainly for studying adsorbed monolayers. Semi- o passive films such as CuO or Cu?0 are often over 100 A 20 thick," and even passive layers such as those found on ° 21 stainless steel are often around 30 A in thickness. In addition, since in most cases the absorption coef- ficient, k, is greater than zero, the thickness, d, of the film must be independently determined before the optical constants n and k can be found. However, this technique does measure the optical constants of the film, which is something differential ref lectometry does not do. Another popular technique is that of electrochemical 72 modulation spectroscopy (EMS)." Here, the modulated parameter is the electrode potential, cell current, or the surface charge of the sample. In its most common form (Figure 3) it is used to measure adsorption on metal 22 Figure 3. Electrochemical modulation apparatus used by Mclntyre.22 The HIS 6070 is a computer used to reduce the data. 2 3 substrates, such as hydrogen on rhodium or iridium, ' or oxygen on gold, platinum, or nickel electrodes. * ' Light from a Xenon arc lamp passes through a monochromator and polarizer and into the optical cell, where it is re- flected from the metal surface and collected outside the cell by a photomultiplier tube. The potential of the sample is controlled by a potentiostat which can be modulated by an external waveform generator. At a signal from an on-line minicomputer, the waveform generator imposes a triangular pulse on the voltage supplied to the sample. The reflectivity as well as the cell current and the electrode potential are digitized and recorded by the minicomputer at 256 time intervals. A number of sweeps (typically 64) are made to increase the signal to noise ratio, and the data is reduced by computer to a plot of -=- versus A. As in the case of differential reflectance spectroscopy, the output is in the form of a series of points rather than a smooth curve. Although electrochemical modulation appears to be an excellent method of studying adsorbed layers, two limita- tions should be noted. First, the reaction must be truly reversible. The formation of an oxide which does not readily dissolve, for example, must be avoided. Second, the reaction must be fast enough to be affected by the 2 4 modulation. (Typical modulation conditions are a sawtooth 9 7 wave, 44 mv peak to peak at 43 Hz." ) For fast, reversible reactions, however, this technique works well. Internal reflection spectroscopy has been used to study adsorption on thin film substrates such as gold28 or 29 platinum. ' In internal reflection spectroscopy, a light beam is internally reflected in a transparent substrate. At the point of reflection, the light beam interacts with the film on the other side of the substrate, and the re- sultant changes in the light beam may be used to charac- 2 Q terize the film. For example, Hansen has observed the o adsorption of molecular dyes on thin (100 A) layers of gold. In addition, electrochemical modulation techniques may be used with internal reflection spectroscopy to study fast, reversible reactions. Such a technique was 2 9 used by Mclntyre and Peck to observe the adsorption of 0 hydrogen on platinum films 100 A thick in .5 M H, SO.. The beauty of the technique is that the observations are made from the electrode side of the interface. Rapid scanning spectroref lectometry is a new develop- ment which permits the continuous measurement of reflec- tion spectra over a wide wavelength range. A recent instrument used by Mclntyre is shown in Figure 4. A rapidly scanning galvanometer mirror (GM) sweeps a high 25 Figure 4. Schematic diagram of a rapid scanning spectro- photometer used by Mclntyre.24 See text for details . 2 6 intensity beam of white light across the surface of a spherical mirror (SM) . The reflected beam is focused by the spherical mirror onto a diffraction grating (GR) at a fixed position, but at a variable angle of incidence. This produces a wide beam of quas i -monochromatic light which is scanned across the exit slit at speeds up to 1000 A/ms . A beam splitter (BS) permits the simultaneous measurement of the incident light intensity via photo- multiplier A (PMT A) and the reflected intensity via photomultiplier B (PMT B) . The signals are sampled in the time domain (which is related to the wavelength by the oscillation frequency of the mirror) by a minicomputer. AR The normalized difference in reflectivity, -5-, of the film free surface (obtained earlier and stored in the computer) and the film covered surface is calculated and displayed. Normally a signal averaging technique is used, wherein typically 512 wavelength scans are used to enhance the signal to noise ratio. Rapid reflectivity changes can be observed, since even with signal averaging a typical scan takes less than one minute. One important advantage of the technique is that surface roughening, such as is possible with electrochemical modulation spectroscopy , is avoided since the electrode potential is not modulated. However, since the scan is so rapid, the 27 high voltage supplied to the photo-multiplier tubes cannot be easily adjusted to maintain the average output voltage at an optimum level. As a result, precision may suffer when the light intensity is low. One final effect common to all these techniques is the possibility of electroref lectance effects complicating the reflectance spectra of the film/metal surface of samples measured in situ. For such samples, the electro- 2 2 reflectance effect can arise from two sources: changes in the dielectric constant of the metal at the surface due to the capacitance of the electrical double layer between the bulk metal and the bulk electrolyte; and, a change in the refractive index of the solution at the ionic double layer. High DC as well as AC potentials can produce these 2 7 effects. The electroref lectance effect can be a valuable probe of the metal surface, or an unfortunate complication of the thin film absorption spectra, de- pending on what is being investigated. The effect does not subtract out of the reflection spectrum when the dif- ference techniques described in this section are employed because the reflectivities of the samples are measured under different conditions. Evidence will be presented in Section III which indicates that the technique of dif- ferential ref lectometry may not produce electroref lectance peaks because the reflectivities of the samples are mea- sured at the same time and under similar conditions. Cuprous Oxide Band Structure Calculations Theoretical studies of the band structure of Cu?0 have been made by Zhilich and Makarov, Elliot, and 32 Dahl and Switendick. The early work of Zhilich and Makarov treated only the r^5 •* r, transition, where T^5 is assumed to be the top of the valence band, and r , the bottom of the conduction band. This is generally agreed to be the fundamental energy gap of Cu20, and gives the oxide its red color. This band gap, calculated according to the method of Kohn and Rostoker was found to be 2.38 eV, in good agreement with the experimentally 30 determined 2.2 eV. 31 Elliot did not calculate transition energies, but instead presented a simple, qualitative band structure derived from symmetry conditions of the cuprous oxide lattice. Both direct and indirect transitions were con- sidered . By far the most ambitious attempt at the calculation of the energy levels of Cu?0 has been by Dahl and 32 Switendick. They calculated the energies of a very 2 9 large number of energy levels, using the augmented plane wave (APW) method proposed by Slater,3 with muffin tin potentials. Their band calculations show the T' + r, transition to have an energy of 1.77 eV, somewhat below the experimental 2.2 eV. Figure 5 shows the calculated band structure. Not all the possible transitions are allowed, and of those which are permitted, many are in- sufficiently intense to be observed. For information on which transitions do occur, one must investigate the optical spectra of Cu20 and attempt to correlate observed structure with possible transitions. The major difficulty here is that while band structure calculations are often qualitatively correct, the assigned energies for favorable transitions are frequently rather imprecise. As mentioned previously, in the calculations of Dahl and Switendick, the r^5 -*• r, transition, which is generally agreed to be the transition defining the fundamental gap in Cu70, is found to have an energy of 1.77 eV. Experimentally, the gap is found to be 2.2 eV — a difference of half an elec- tron volt. Hence while band structure calculations can provide a useful qualitative picture of the structure of a thin film differential ref lectogram, quantitative data is best obtained from optical spectroscopy. 30 \_ i. V N 1_ \ TA R MITAX Figure 5. Band structure diagram for cuprous oxide (CU2O) calculated by Dahl and Switendick . 32 51 Optical Spectroscopy 35 Brahms and Nikitine have measured the absorption and reflectance spectra for single crystals of Cu90 in the 2.5-6.5 eV region. Absorption was measured at 77 K and 295 K, and reflection at 77 K. A good deal of peak sharpening occurs in the absorption spectra at low tem- peratures. The correlation between peak positions for absorption spectra and reflection spectra as shown in Table 1 is quite good, as it must be. The authors associate the transitions at 2.58, 2.62, 2.71, 2.74, and 4.33 eV with exiton formation. The strong peak at 5.5 2 eV is attributed to a maximum in the joint density of states . Balkanski, Petroff, and Trivich have measured the reflectivity of Cu20 in the ultraviolet (2-25 eV) at 300 K, and the reflectivity in the visible (200-500 nm) at 30 K. They then use the Kramers -Kronig relationship to determine e- for the 2.5-6 eV region of the spectrum at 30 K. The reflectivity data for Cuo0 at 30 K corresponds L 35 well to the earlier work of Brahms and Nikitine. Both mechanically polished as well as polished and etched samples were measured, as shown in Figure 6. It is evi- dent that most of the structure, including all of the exiton peaks, is dependent upon the surface state of the 32 Table 1 Comparison of Peak Positions for Absorption and Reflection Spectra^S Absorption Reflectivity Peaks [eV] Peaks [eV] 2.61 2.58 2.62 2.71 2. 71 2. 75 3.62 3.62 3.9 4. 33 4. 33 4.48 4.48 4. 74 4. 74 4.86 4.86 4.94 4.94 5.08 5. 36 5. 39 5.52 5.53 5. 79 5.84 6.45 33 3500 I M(cm')i_ Figure 8. Electroref lectance spectrum of CU2O at 85 K and no DC bias. The modulating electric field was 60 KV/cm.42 39 X-ray investigations showed that in the early stages of oxidation the crystals produced were amorphous. This implies that thermally grown oxides may not have a well- defined crystal structure and hence not a well-defined band structure during early stages of growth. In most of these cases of thermally grown Cu?0, the aim was to produce bulk samples, rather than thin films. Electrochemical methods are probably better suited to growing thin films on copper substrates than thermal methods, since electrochemical methods tend to produce films which are more dense and compact. Characterization of the film is, as usual, the major problem. The growth of oxide films of spherical single crystals of copper in distilled water was studied by 20 45 46 Kruger." ' '' Using a special oscillating crystal technique he was able to establish that Cu„0 was the film formed on copper in distilled water exposed to air. This is the same oxide which is slowly formed when copper is exposed to air. Further, when pure oxygen was bubbled through the water, the film initially produced was Cu?0. This is not unreasonable since any CuO which begins to form would tend to be reduced to Cu20 by the copper sub- strate. These Cu20 films were observed to form very rapidly and reach a limiting thickness after about two 4 0 hours. The precise film thickness, as measured by standard ellipse-metric techniques, varied with the crystallographic orientation of the copper, but averaged o about 90 A for this case. After long times a porous, loosely adherent CuO layer was observed to form on the Cu20 layer. The formation of the CuO overlayer apparently did not prevent the continued growth of the Cu?0 under- layer. One variable was found to influence the growth of Cu^O, however: light. Cuprous oxide is a p-type semiconductor. Since its band gap of 2.2 eV is in the visible part of the electro- magnetic spectrum (about 560 nm, or yellow- green) it is not unexpected that light might have some effect on its electronic properties. That these electronic properties affect the kinetics of growth is somewhat more surprising. When Cu20 films are illuminated by the appropriate wave- length of light, electrons are transferred from the valence band to the conduction band, where they can mi- grate to the copper electrode. This process inhibits the growth of the Cu20 film by a process not completely under- stood. Whatever the exact mechanism may be, inhibition of Cu20 growth has been observed by Kruger on single crystals 45 46 of copper. ' Similar results were obtained by Ives and 47 Rawson for polycrys talline copper. Kruger 's work with 41 single crystals showed a decrease in Cu?0 film thickness of about one-third for samples illuminated by a 500 watt tungsten light over similar samples which were not illuminated. No effect on the film growth rate was ob- served if the distilled water was kept saturated with oxygen. Attempts to employ a monochromator with the light source to determine the spectral response of the inhibition of film growth were not successful due to the lack of light intensity. This implies that monitoring film growth with the differential ref lectometer or an ellipsometer probably would not inhibit the growth of a Cu^O film. Summary Of the several methods related to differential re- flectometry which may be used to investigate thin film corrosion products, ellipsometry is the most common. In its basic form, ellipsometry can determine the optical constants n and k to a high degree of precision if the thickness of the film is known. Recent advances have eliminated even that requirement, provided the absolute o reflectivity can be measured, or if a reasonable guess 9 10 can be made as to the thickness of the film. ' Very precise values of n, k, and d may thus be determined from 42 a classical, well-established instrument. The only re- maining drawback is the slow speed of the classical ellipsometer , and even this barrier is being overcome through the use of automatic and modulated ellipsometers . For thin adsorbed films, many spectroscopic techni- ques are available. Differential reflection spectro- 17-19 . scopy is very similar to differential ref lectometry , but employs polarized light and measures the reflectivity before and after film formation rather than concurrently. Although it can determine the optical constants, it can only do so for very thin monolayer films. For the study of fast, reversible reactions in situ, electrochemical modulation ' " is ideal. Here an adsorbed film is induced to form and redissolve by modu- lations of the electrode potential. Its only real dis- advantage is that the number of possible applications is small. Also somewhat limited in application is internal reflection spectroscopy, which has the unique advantage of observing film formation from the electrode side of the interface. Its chief limitation is that light must pass through the electrode. Finally, for very fast reactions or transient phenomena, rapid scanning spectroref lectivity can be a useful method of measuring changes in reflectivi- ty. 4 3 Some such techniques have the advantage over dif- ferential reflectometry in that they can determine the optical constants of the thin film corrosion product. Some, such as electrochemical modulation or rapid scanning spectroreflectivity, can study very rapid reactions. However, all except ellipsometry are more or less limited to very thin (monolayer) films. Such films are important in studies of catalysis and adsorption, hut passive and semipassive layers are too thick to be analyzed for n and k using these techniques. Differential reflectometry is well-suited to the study of these films in the 5 A-500 A range, but requires some way other than the optical con- stants to identify the film. Several avenues are open to one wishing to identify thin film corrosion products with differential reflecto- metry. Some, such as x-ray or various forms of electron spectroscopy, were discussed in Section I. Other methods include the correlation of peak position with either band structure data, or reflectivity spectra. The most complete band structure study of Cu?0 is that of Dahl and Switendick. " While it agrees well qualitatively with observed structure in the reflectivity spectrum, the energies of transitions do not. More de- tailed calculations are required before band structure 44 data alone may be used to identify corrosion products without additional data from reflectivity curves. Several techniques have been used to record the optical spectra of Cu20. Most agree on the major struc- ture, but various techniques differ in their sensitivity to exiton structure and to complications from electro- reflectance effects. The data of Balkanski et al . is most important since it shows that structure due to exiton formation may not be visible in unetched specimens. Table 2 summarizes the reflectivity data and band structure calculations discussed in this section. Only room temperature data are shown . The calculations of Dahl and Switendick are shown shifted so that the fundamental gap occurs at the experimental energy, with the original data in parentheses. This translation may not be correct, but probably results in more accurate values than the original calculation. It is given here as an approxima- tion only. Agreement between the different experimental techniques is good, which implies that peaks at similar positions should be visible in differential ref lectograms of Cuo0. 4 5 Table 2 Major Peak Positions in C112O Optical Spectra Compared to Transitions Calculated from the Band Diagram (see text) Reflect ivity 36 e2 Electrore Ref. 40 f lectance Ref. 4 2 3? Calculated 2. 5 2.7 - 2. 59 2.73 2.25 (1.77) 3.4 - 3.35 3.60 3.66 3.77 3.84 (2.86) 3.7 3. 7 3. 78 3.95 - 3.74 (3.26) 4. 3 4.4 4.08 4.25 4. 35 4.90 4.41 (3.63) 4. 7 4. 8 - 4. 72 4.76 4.92 4.83 (4.35) 5.0 5.3 - 5.03 5.21 5.07 (4.59) SECTION III EXPERIMENTAL PROCEDURE Introduction Any analytical technique must be capable of consis- tent results. Frequently this is the most difficult aspect of the development of a new technique, since un- expected variables in the experimental process may produce inconsistent results. The application of differential ref lectometry to corrosion studies is particularly sus- ceptible to this problem for three reasons— it involves thin films, electrochemistry, and a new analytical techni- que . One reason thin films of corrosion products are difficult to study is because they are difficult to de- tect. They are known to be delicate, often easily al- tered, and frequently nonuniform, either across the specimen or in depth profile. Electrochemical techniques, being thermodynamic in nature, are easily influenced by small changes in experimental conditions. For example, unexpected temperature changes or power interruptions may produce anomalous results. Finally, the analysis techni- que, differential ref lectometry , while well-established 4 6 47 for compositional modulation studies of alloys, " has never been applied to electrochemical studies. It is therefore of the utmost importance that the sample preparation and experimental procedure be carried out in a careful and consistent manner, in order that the results be as reproducible as possible. Sample Preparation Metallographic Preparation Copper samples were prepared from a 3/4" (19 mm) diameter bar of 99.999% pure copper.* Each sample was lathe cut from the bar to form a dish 3/4" (19 mm) in diameter and approximately 3/8" (9 . 5 mm) thick . In order to eliminate any possible contamination by the cutting operation, each dish was immersed in 20% nitric acid for fifteen hours. A small, 1/16" (1.6 mm), hole was drilled in the back of the sample to a depth of about 3/16" (4.8 mm) to facilitate the attachment of the electrode wire . Each sample was then ground flat with a succession of 180, 320, and, finally, 600 grit silicon carbide papers. Grinding was accomplished on a metallographic *0btained from the A. D. McKay Corp., New York, New York. 4 8 polishing wheel while the papers were lubricated with liquid soap and water. Due to the fact that the silicon carbide particles tend to become embedded in soft mate- rials such as copper, the minimum amount of polishing necessary to properly finish the specimen was performed. Since it was found that even these measures did not eliminate silicon carbide embedding, further polishing was performed using Microcut paper.* Microcut paper is a mildly abrasive , napped paper which produces a surface finish similar to 600 grit silicon carbide paper, but contains no silicon carbide or other particulate abra- sives. The cutting action is a product of the nap of the cloth alone. By polishing each sample for a sufficiently long period of time on several pieces of Microcut paper lubricated with liquid soap and water, it was found that virtually all silicon carbide contamination could be eliminated, as confirmed by microscopic examination. Fortunately such elaborate measures were found to be necessary only during the first polishing of the as-cut specimens. Used specimens with mild amounts of corrosion in the form of thin film layers could be cleaned with Microcut paper alone. As mentioned above, this produced 'Microcut paper, trademark of Buehler, Ltd., Evanston, I 1 lino is 4 9 a sample with a finish similar to that produced with 600 grit silicon carbide paper, but without the silicon car- bide contaminants. Such a finish completely removed all traces of the thin films studied in this work. After polishing with Microcut paper, each sample was fine-polished with 6 micron followed by 1 micron diamond paste on a felt cloth using Metadi* fluid as a lubricant. This polishing technique produced a uniform, mirror-like finish on the sample. In order to fasten the sample wire to the specimen, a tiny brass fitting was constructed. The fitting was in the form of a truncated cone with a maximum diameter of about 1/16" (1.6 mm). The sample wire, which was insu- lated with Teflon** to ensure freedom from attack by the electrolyte, was soldered to the end of the fitting. This assembly could be attached to the sample by means of a press fit. Finally, the fitting and most of the sample were covered with a special lacquer (Miccrof lex***) which previous experience had indicated would not be affected by ^Metadi fluid, trademark of Buehler, Ltd., Evanston, 111 inois **Teflon, trademark of E. I. DuPont , Wilmington, Delaware ***Miccroflex, trademark of Michigan Chrome and Chemical Co., Detroit, Michigan 50 the electrolyte and would prevent the covered section of the sample from corroding. Special Procedures for In Situ Samples Samples prepared for in situ analysis require an optically transparent layer for the insulating coating, since the light beam must be able to see a corroded and an uncorroded part of the sample at the same time while the sample is in the electrolyte. It was eventually established that collodion provided the most satisfactory coating . From a theoretical point of view, quartz would seem to provide the best coating. It is optically transparent, insulating, and very resistant to chemical attack. How- ever, quartz cannot be easily deposited on the copper surface. Extensive attempts to sputter quartz on the sample face were not successful due to the nature of the sputtering process. When quartz, SiO?, is sputtered off a quartz target, SiO- particles are not the only species produced. The momentum transfer process which makes sputtering possible partially dissociates the SiO? into Si, 0, 0?, and SiO as well as the predominant Si02- The resulting free oxygen in the chamber reacts with the copper sample before a protective layer of SiO., can be 51 deposited. The result is oxidized copper, covered with an insulating SiCL layer. The effect is enhanced by the high temperature of the plasma and the high energies imparted to the oxygen ions by the sputtering action. In addition, in order to obtain SiO~ layers free from SiO contamina- tion, the manufacturer* suggests adding additional oxygen to the system and sputtering in an oxygen enriched argon atmosphere. This only provides more oxygen to oxidize the copper before a protective layer is formed. (The situa- tion is no problem to the major user of quartz sputtering equipment, the semiconductor industry, since they are generally sputtering onto silicon.) Silicon monoxide, SiO, was another unsuccessful can- didate. Although it can be evaporated instead of sput- tered, it is not entirely optically transparent. A SiO coating has a yellow-brown tint which is readily detected by the differential ref lectometer . Attempting to evaporate SiO in an oxygen enriched atmosphere and deposit Si02 on the sample produces only the oxidation and de- struction of the boat used to evaporate the SiO. Evidently some kind of mechanically applied chemical compound similar to Miccroflex but transparent was *Perkin-Elmer Co., Palo Alto, California 5 2 necessary. Miccroflex may be thinned out to a great extent and still remain protective, but it also remains purple. The color may not be apparent to the eye, but the differential ref lectometer shows it quite clearly. Other commercial compounds used in electroplating such as Miccroshield* showed similar results. It is known that Tygon tubing, noted for its chemical resistance, can be liquified by the chemical Tetra- hydrafluoran (THF) . This can then be readily applied to specimens by a dip method, and subsequent air drying restores the physical and chemical properties of the original Tygon. Transmissivity studies showed it to be transparent from 800 to approximately 350 nm, which en- compasses much of the present range of the instrument. Unfortunately, the dissolved Tygon would not dry to a uniform thickness. This produced interference colors on the film which completely obscurred any differential re- flectometry data. Attempts to produce films of Tygon o thinner than about 2000 A in order to eliminate the inter- ference effect produced films with no appreciable degree of adherence to the copper sample. This led to a small ''Miccroshield, trademark of Michigan Chrome and Chemical Co., Detroit, Michigan 53 gap between the copper and the "protective" film, causing massive crevice corrosion. Thus far the most satisfactory coating has been found to be collodion, a waxy organic compound used to make replica specimens for transmission electron microscopy. It is available dissolved in either amyl acetate or alcohol and may be thinned to any desired degree by the addition of additional solvent. Since alcohol of extreme- ly high purity is easily obtained, and has the advantage of rapid drying time, it was chosen over amyl acetate as the solvent for this work. Collodion layers may be deposited on half the sample by the simple method of dipping half the sample in the dissolved collodion. The collodion then dries to a uniformly thick, fairly adherent layer in about thirty seconds. The physical properties of the collodion layer are similar to the Tygon coatings. Thick layers will produce interference colors due to non-uniformity of the film. Extremely thin layers have little adherence to the copper sample, and produce crevice corrosion. The advantage of collodion over Tygon is that the collodion is more readily dissolved, dries faster, and produces more uniform films. Even so, a film sufficiently thin to minimize absorption of the light beam and eliminate interference 5 4 colors will break down or loose adherence after twenty- four to forty-eight hours in the electrolyte. Further research in this area of transparent, insulating coatings is clearly needed. Electrolyte Solutions Electrolyte solutions were prepared from reagent grade chemicals and distilled water which had been passed through two ion exchange columns. Typical resistivity was on the order of 10 ohms-cm. Nil chloride solutions were prepared according to Table 3. Immediately before use electrolytes were deoxygenated by vacuum deaeration. All beakers, flasks, and corrosion cells used for nil chloride work were used only for nil chloride work, and were periodically cleaned with nitric acid to ensure freedom from chloride contaminants. The Differential Ref lectometer Introduction The differential ref lectometer is a device which measures the difference in reflectivity of two samples as a function of wavelength. The technique is a form of modulation spectroscopy where the modulated parameter is the surface composition of the samples. Originally de- 55 Table 3 Composition of Electrolyte Used in Nil Chloride Experiments pH Buffer Na OH 4 100 ml 1.0 M KHP 5 100 ml 1.0 M KHP 108 ml .5 M 6 100 ml 1.0 M KHP 180 ml .5 M 7 14 3 ml 0.7 M H BO 10 ml .1 M dilute to 8 14 3 ml 0.7 M H^BO 100 ml .1 M 2000 ml 9 14 3 ml 0.7 M H BO 100 ml .5 M 10 14 3 ml 0.7 M H BO 184 ml .5 M 11 132 ml 0.7 M H BO 220 ml .5 M 56 veloped to monitor changes in the band structure of metals with alloying, it is used here to investigate the growth of thin films of corrosion products on metal substrates. In this case the differential ref lectometer scans a light beam between two areas of the same sample, one of which is clean and the other covered with a thin film of corrosion product. The reflectivity of the copper substrate, which is common to both sides of the sample, does not contribute to the differential ref lectogram, while the absorption spectrum of the corrosion product does contribute. The Instrument A schematic diagram of .the instrument used in this work is shown in Figure 9. Light from a high pressure Xenon light source* is passed through a scanning mono- chromator** and reflected by a flat mirror onto an oscillating mirror. The oscillating mirror scans the light beam across the two areas of the sample at a fre- quency of 60 Hz. The size of the area to be scanned can be varied by changing the voltage applied to the mirror. *Xenon light source, Mc Pherson model 613, 75 W, Mc Pherson, Inc., Acton, Massachusetts **Grating monochromator , Mc Pherson model 218, Mc Pherson, Inc., Acton, Massachusetts 57 - .* i '4/ light from the monochromator must be blocked except when actual measurements are being made. Summary The differential ref lectometer was used to measure the difference in reflectivity between the corroded and uncorroded areas of electrochemical specimens. The speci- mens were prepared using standard metallographic techni- ques, with extra care being used to eliminate carbide em- bedding and sample contamination. Electrolyte solutions 71 were prepared from reagent grade chemicals and deaerated before use. Normal as well as in situ corrosion cells were designed to have only inert substances in contact with the electrolyte. A plot of the normalized difference in reflectivity of the clean half of the sample minus the corroded half can be obtained as a function of wavelength over the spectral range 800-200 nm. SECTION IV RESULTS AND DISCUSSION Introduction The purpose of this work was to investigate the pos- sibility of using differential ref lectometry to identify and study thin film corrosion products on metal sub- strates. For this initial investigation, copper was chosen as the base metal because it has easily formed thin film corrosion products which are visible to the eye This ensures that there are optical transitions which would be detected by the differential ref lectometer , since the instrument can examine a wavelength range greater than that visible to the eye. In addition, work had previously been performed here on copper alloys using the differential ref lectometer , so the mechanics of pre- paring samples were known as well as the theory of inter- band transitions in copper and their relation to peaks in differential ref lectograms . 52 Both the theoretical or equilibrium (Figure 12A) 5 3 and experimental (Figure 12B) Pourbaix diagrams are available for copper in nil chloride solutions. In the basic half of the diagram (pH 7-14) three species are 72 73 X CL — • o o | O Q) m o a o o o o r 3 rO u CD XJ +-> P) • H O O <* X -d h (D 3 h o rt X Ph rci e H Mld o CN +-> OJ u 1 o v — ' l—l Vj ft a; rt u H W> i— i tf rt •H ■H T3 +-> c X o • H 'm nj o H-t CVJ o c o ■ H W) 0) O !-. o o rO 3 u c/i 0) 3 U O cU X P. CD 00 in a ct3 - W m u o CO e > in o o O + t— I CD rt rt ■H •H TJ 4-> fi X CD •H S-h Oj CD Xi m fH m 3 •H o Q &. 77 °< P-CTi *• — " £*& s< LJ 1/)00 s c o o • H +-> • 3 rt o Sh !/) O id +-> 'H U rH r-H £ rt 3 ■H c ■P ■ H c £ 0 3 in r-H 0) cti 4h m o\° ■H 1-1 4-> rt c o r-H •H +-> rC cd +-> r-H •H 3 £ T3 O H e rH „ rt rH u GJ H Ph Ph ft perforin these calculations, the index of refraction, n, and the index of absorption, k, for both the film and the metal substrate must be known as a function of wavelength. The equations necessary to calculate the reflectivity of the film covered metal surface have been described by 7 Heavens. Referring to Figure 16, n„ is the index of refraction of the medium surrounding the sample, which here is either air (nQ = 1.00) or electrolyte (n„ = 1.33). The thin film corrosion product has optical constants n, and k,, and the thickness d, while the metal substrate has optical constants n? and k? . We have, then, _ no2-nrki _ 2noki ' 2 — 2 1 ~ 2 7 (nQ+n1) +k1 (n0 + IV +kl 2 2 2 2 n"-n?+k, -k„ 2 (n, k? -n?k, ) 2 " (n^n^^Ck^k^2 2 " (n1 + n2)2+(k1 + k2)2 2Trk,d 2Trn,d a = — t — [radians] y = - — t [radians] The reflectivity, R- , of the film/metal surface is (gx + h2)e"2fX + (g2 + h2)e2a + Acos2y + Bsin2y 2 e2ci + (g2+h2) (g2 + h2)e"2a + Ccos2y + Dsin2y where A = 2(g1g2+h1h2) B = 2(g1h2-g2h1) C = 2(g1g2-h1h2) D = 2(g1h2+g2h1) 79 n0 (AIR OR WATER) Figure 16. Schematic representation of a typical sample, with an oxide film covering half the exposed surface . 8 0 The reflectivity of the clean copper surface is much easier to calculate. Calling R, the reflectivity of the copper surface and using the same notation for the optical constants , 2 2 (n?-n ) +k: Rl = % 2 & (n2+n0)-+k2 The normalized difference in reflectivity is easily obtained : AR R1~R2 ~ r~Tr — (63 R ' 1 K2, t— 2 — ] These equations should reproduce the experimental differential ref lectograms for non-in situ experiments if nQ = 1.00 and if values of n and k are available in the literature for copper and its oxide as a function of wave- length. In addition, if the effect of collodion is neglected, it should be possible to reproduce differential ref lectograms of in situ experiments. In this case, nQ = 1.33, the index of refraction of water, which is approximately that of the electrolyte. These calculations have been performed for the case of a cupric oxide (CuO) film since the optical constants for that compound have been published as a function of wavelength. The optical constants for copper which were used in these calculations are those of Schultz and 54 55 Tangherlini ' and are given in Table 4. Figure 17 is the plot of the variation in AR/R for CuO film on copper substrates as a function of film thickness for a fixed wavelength of light. The particular wavelength used in these calculations was 600 nm, but the curve is similar to those calculated for other wave- lengths. The curve shows a large increase in the dif- ferential reflectivity for the first few hundred angstroms of film thickness, followed by strong oscillations which are gradually damped out. The oscillations are caused by interference effects which occur when the film thickness is of the same order of magnitude as the wavelength of the incoming light. For constructive interference in this case, the film thickness must be d = Ji- (7) 2n, *- ■* where n, is the index of refraction of the film, as be- fore, and m is the order of the interference. The damping effect is due to the fact that as the film thickness, d, increases, less light penetrates to the underlying metal surface. The reflectivity of the film/metal surface slowly approaches the reflectivity of a bulk specimen with the composition of the film. Further increases in the 82 c o e e i— i c ■ H 4h o o O vO 3 U 0 £ oj o h 4-> O 4h TJ o tfl e 00 3 4h M O U; •H g r-i o •h 4^ +-) o u G -G 3 +-> 4-1 Cft G 03 o r-H (/) ID cd > 01 Iff! V Pi 0) < X H C ■H G 0 o +J > V) 8 3 Table 4 Optical Constants n = n+ik for Copper and Cupric Oxide (CuO) as a Function of Wavelength^ , 54 , 55 A [nm] Co] n jper k Ci n lO k 450 .87 2. 20 2.45 .744 500 .88 2.42 2.59 .650 550 .72 2.42 2.57 . 539 600 .17 3.07 2.60 .449 650 .13 3.65 2.65 .345 700 .12 4.17 2.64 .236 750 .12 4.62 2.65 .170 800 .12 5.07 2.62 .132 84 thickness of the film do not change the reflectivity. Since the clean copper surface has a constant reflectivity at a given wavelength, the differential reflectivity does not change after the film reaches a certain minimum thickness. For the case of the CuO film this occurs at o roughly 7500 A, beyond which the differential reflectivity is a constant 121% . Other wavelengths of incident light show similar oscillations with slightly different periodicity. This interference can produce peaks in differential reflecto- grams which are not due to interband transitions but rather are due to the thickness and index of refraction of the film. As a result, differential ref lectometer spectra of thin films on metal substrates can have struc- ture which varies as the film thickness changes. For example, Figure 18 shows the theoretically calculated o differential reflectogram for a CuO film of 100 A thick- ness. Two curves are shown, one for which the medium surrounding the sample is air (nn = 1.00) and one for which the surrounding medium is assumed to be electrolyte (n~ = 1.33). Both curves are essentially the same, indi- cating that there should be little difference between spectra taken in situ or non-in situ if the effect of collodion is neglected. These curves differ slightly from 8 5 u • H ■M O U q-i o £ d so o u c e o g M cu o <-H o u^ *o •H "3 ii T3 ^< 0) +J (J D (U fH +-> ■P rt CO ^ G +-J o CO U JD 3 X to t— ( 1— 1 Jh Cfl o a; u ^ o rt a> -G G H ) 86 the theoretical curve calculated for a 35 A thick CuO film in air (Figure 19) but are significantly different com- o pared to the spectra of a 500 A film (Figure 20) . Finally, Figure 21 shows the calculated differential re- o flectograms for a 7500 A CuO film surrounded by air or by o an electrolyte. Film thicknesses greater than 7500 A should produce spectra similar to Figure 21, since in this case the light beam will not be able to penetrate the film all the way to the film/metal interface. As can be seen by comparing Figure 21 with Figure 19, such curves bear no resemblance to those of very thin films. Fortunately, as will be described below, the thin film corrosion products which were generated in this work have a thickness which appears to be significantly smaller than the critical thickness at which interference effects begin to distort the differential ref lectograms . o Using Equation 7 with n-, = 2.1 and A = 2500 A, the o first interference maximum is at about d = 600 A. Wave- o lengths greater than 2500 A (which is the wavelength where most differential ref lectograms begin) place the first maxima at higher thicknesses. Therefore for thicknesses o up to at least 200 A the differential reflectogram should u o< £ o g W) o +-> u o 0) o "-H O 4-1 O +-> cd IT) ^ c +-> o CO o rQ 3 X t/1 o •H P. o 4-> O CD o rt TO t/) >H Pi +-> o t/1 U ,Q 3 X LO I— ( 1— 1 ^ OJ o 0) u fH o oj 0; Xi c H o 8 9 180 400 Figure 21. Theoretically constructed differential reflec togram of a CuO film 7500 A thick on a copper substrate. A = 600 nm. 90 not show an appreciable effect from interference. Film o thicknesses greater than 7500 A should also be free of interference effects, since the light beam should not reach the copper substrate under the corrosion product film. Thus there may be defined three distinct ranges relating film thickness to differential ref lectometry . In the first range, where film thicknesses range from 0 to at o least 200 A, the structure in a differential reflectogram is constant and unique for each kind of film. The thick- ness of the film determines the intensity of the dif- ferential reflectivity. In the second range, extending to o a thickness of about 7500 A, interference effects dominate the spectra. Finally, in the third range, where the film o thickness is greater than 7500 A, the differential re- flectogram will again be unique for each corrosion pro- duct. However, peak intensities will be constant and unaffected by increasing film thickness. The differential ref lectometer should be most useful in the first range, where intensity is related to film thickness. Figure 22 shows the variation in AR/R with film thickness for the same conditions as Figure 17 (CuO film, copper substrate and reference, nfl = 1.00), but over o the more restricted film thickness range d = 0 to 200 A. This curve may be fitted to either a power curve, 91 ^ T3 CD (D Oh e PhS 0 w 0 i/i o o 03 (/) CVJ C, -H O 1— 1 bO •H -H ■4-1 rH O <+-! 3 O u n3 +-> O O CD m CO Oj a M CD ox • H H X 4-> 1 1 o< rH < » •H O o (J) <-H O CM o to 0 0 UJ +j 2: a 0 0 *: •H 0 ■M II CJ X 3 H 4h cd 03 pj > 03 _l L0 rH 03 o m Li_ ariation in AR/R ubstrate over the 0 be 600 nm. o o o o o 00 o o o OJ > rt rt -M sh x e Cfl CL,i-H O -H •M 4-> 4H U rt <^i o •H O ■M O S C« ni O + ^H ^ &0 o rt CD •p ^ £ u rt o Ch 6 u o 94 more like those of copper at the copper/film interface, and more like those of bulk CuO at the film/electrolyte interface. As a result, calculated differential reflec- tograms prepared from the optical constants of thermally grown compounds should not be expected to exactly match the differential ref lectograms of electrochemically pre- pared compounds. o A CuO film 25 A thick may seem surprisingly thin, but films near this thickness are necessary to give good agreement between the experimentally observed and the theoretically calculated differential ref lectograms . In addition, Kruger observed CuO and Cu?0 film thicknesses o on the order of 100 A or less for films grown in air or water. In addition, measurements were performed on several films using an interf erometric microscope.* o Precise measurements of film thicknesses less than 100 A are not possible with this instrument, so an exact mea- surement of the film thickness was not obtained. How- ever, the instrument did unambiguously indicate a film o thickness of less than 100 A for each film measured. In order to be observed with the interferometric microscope *Model M-lll Angstrommeter, Sloan Instrument Co., Santa Clara, California 95 the film must be overcoated with a thin layer of eva- porated aluminum or other metal. This involves placing the sample in a vacuum which could conceivably have altered the film, producing inaccurate measurements of the thickness. For this reason measurements with an inter- ferometric microscope are not entirely reliable for corrosion applications. Finally, a working ellipsometer has become available recently at the University of Florida. Measurements of a typical CuO film, prepared by potentiostating in a nil chloride solution of pH 9.2 for eighteen hours at +100 mv o SCE, indicate a film thickness of about 45 A. All these measurements point to film thicknesses for typical speci- o mens of well less than 200 A. In this thickness range, the peak positions in the differential ref lectograms remain essentially constant, and the peak intensities may be used to calculate the film thicknesses to a high degree of precision. Non- in Situ Experiments In this section the results of experiments which were not performed in the in situ cell will be discussed. Longer exposures to the electrolyte are possible using this method than in true in situ studies because of the 96 tendency of the protective collodion film, used in in situ experiments to cover the copper reference side of the sample, to break down after a certain amount of time in the electrolyte. The disadvantage is that changes in some films are possible between the time of removal of the sample from the electrolyte and the measurement of the differential ref lectogram. Three types of film formation were studied: films formed in air; films formed in dis- tilled water; and, films formed in electrolyte solutions during potentiostat ing . Films formed during exposure to the atmosphere were prepared by the simple technique of covering half the sample with a protective layer of Miccroflex and per- mitting the sample to remain exposed to the atmosphere. An inverted beaker or glass plate was used to prevent dust from settling on the specimen. Similar experiments were performed with samples exposed to an atmosphere of pure oxygen. In these cases the sample was placed in an airtight dessicator which was repeatedly flushed with oxygen . The results of the experiments where the sample was exposed to the uncontrolled atmosphere were seldom com- pletely reproducible. Greater or lesser amounts of various oxides could be formed depending on the initial 9 7 conditions, length of exposure time, and relative humidi- ty. As a result, between ten and twenty runs were made for each variation in the experiment, under conditions which were kept as identical as possible, in order to determine trends of film formation. Normally approximate- ly 80% of the runs produced essentially identical results. Samples exposed to a normal indoor atmosphere for a moderate period of time (one to three days) yielded dif- ferential ref lectograms essentially the same as Figure 13; that is, indicative of cuprous oxide, Cu?0. Other samples exposed to air for longer times (five to fourteen days) tended to produce CuO patterns, as did all the samples exposed to a pure oxygen atmosphere. Kruger has pointed out that Cu?0 is the oxide produced upon initial exposure of copper to an uncontrolled atmosphere. However, the 5 7 Cu-0 phase diagram predicts that CuO is the stable species if copper is exposed to dry air below 375°C. This apparent inconsistency can probably be explained by the fact that the normal Florida atmosphere is not at all dry. Water vapor probably induces the formation of Cu?0 rather than CuO. In addition, CuO tends to be reduced to Cu90 in the presence of metallic copper, so it is likely that there is at least a monolayer of Cu-0 at the inter- face between Cu?0 and Cu. Hence copper exposed to the 98 atmosphere for a brief period of time may well show more of a Cu?0 pattern than a CuO pattern. Presumably exposure to the atmosphere for longer times or exposure to pure, dry oxygen would favor the growth of the equilibrium oxide, CuO, as was observed. In order to study the films formed in distilled water, samples were prepared in the usual manner, par- tially covered with Miccroflex, and immersed in a beaker containing approximately 500 ml of triply distilled water. A magnetic stirrer was used to maintain circulation. In some cases the water was simply exposed to the atmosphere; in others oxygen was bubbled through the water at a rate of about 10 cc/sec. Samples placed in distilled water which was exposed to the atmosphere usually produced differential reflecto- grams indicative of mixed oxides (CuO and Cu?0) with the Cu20 being predominant. In a few cases only Cu^O was apparent. This is in agreement with the work of Kruger, who also found Cu^O films. Kruger has also noted that the Pourbaix diagram predicts the formation of CuO, rather than Cu20, under these conditions. One of the advantages of the differential reflecto- meter is its ability to clearly demonstrate the presence of mixed oxides, such as are found in the above experi- 99 merit. The results seem to show a simple addition of the spectrum of one compound to the spectrum of the other. In the case of samples placed in distilled water through which oxygen was continuously bubbled, mixed oxides of Cu20 and CuO were again produced. A typical result is shown in Figure 24, where the structure at 310, 330, and 380 nm is characteristic of Cu~0 (see Figure 13), and the rising tail visible below 300 nm is characteristic of CuO (see Figure 14) . Although a quantitative analysis of such mixed oxides is not possible at this time due to a lack of information on the optical constants of the Cu?0 phase and to the increased mathematical complexities of multilayer reflection, such analyses are certainly pos- sible in principle. The most extensive studies to date have been in boric acid buffered electrolyte with samples potentiostated in some particular region of the Pourbaix diagram. For the most part these experiments agree well with the Pourbaix diagram determined experimentally by electrochemical techniques (Figure 12B). However, the differential ref lectometer provides additional insight into the com- position, thickness, and uniformity of the corrosion pro- ducts. An example of the additional information which may be obtained using optical techniques is shown in Fig- 100 p CD p nj £ o xi o o o o 00 < cr O ro cfl (J E p P -H bA-H O £ CD +-> P U « 3 CD CO 4-> rH P U CD O P P X -P C/) rH £ 03 CD CD •H CD 4=1 P P H CD U) P -H • CD CD p Mh CD

•H O 03 101 ure 25, which shows a differential ref lectogram of a sample potentiostated at -50 mv SCE and pH 9.2 for eighteen hours. Three scans are shown, taken at three different areas of the sample. Several points are readily apparent. 1) The thin film corrosion product is primarily Cu20, as indicated by the structure at 315, 330, and 380 nm. 2) The rising slope between 300 and 200 nm indi- cates a small amount of CuO present in the sample. 3) Variations in intensity indicate differences in film thickness, with the film thickness decreasing from position C to position A. Mixed oxide films such as those indicated by the spectra of Figure 25 are not uncommon. This sample is typical of those potentiostated near the Cu?0/CuO equili- brium line (-50 mv SCE at pH 9.2 is essentially on the equilibrium line) . It illustrates the fact that in such regions of the Pourbaix diagram one corrosion product is not much more stable than the other. Initial conditions, such as the presence or absence of a spot of oxide, or a local homogeneity in the copper or electrolyte, may in- fluence which corrosion product may form in any specific microscopic area. Samples potentiostated well into a 102 3 rt -w CD u i-H CO a e > co 6 V) o 0) LO E i rt v— ( CO CD a> C rC -H 4-> rH m 6 O 3 ■H I/) Sh CO ^ CD -H E (h i-H CO -H 3 c 3 u <+H v-s o o (NI CO 3 6 U CO Jh CD MX O +-> +-) U fH CD cO rH CD m C CD fmj i/i CD U H(J 3 cO cO O •H +J ,C +-> CO rt o c CD -i-l CD ^ +J CD CD id +J <+H CD X m +-> to •H O -H Q ftd) 103 specific region of the Pourbaix diagram consistently show only the expected corrosion product, as in the case of the differential ref lectograms shown in Figure 13 (-225 mv SCE, pH 9.2), which is typical of a pure Cu~0 film, and Figure 14 (+200 mv SCE, pH 9.2), which is typical of a CuO film. A large amount of data was collected with the aim of defining more precisely the positions of the lines de- lineating the areas of stability of the experimental copper/nil chloride Pourbaix diagram. This is complicated by the fact that these lines are coexistance lines, so that both compounds are present near the boundaries, as confirmed by the differential ref lectometer . Furthermore, the boundaries can be slightly time-dependent. For example, a sample potentiostated just into the CuO region of the Pourbaix diagram, but near the Cu^O area, might show Cu^O components in the structure of the differential reflectogram if the exposure time was brief. In this case the CuO which formed against the copper surface might be reduced when the potential was removed. Alternately, as was mentioned earlier, a thin layer of Cu20 might always be present at the Cu/CuO interface, regardless of where in the Pourbaix diagram the experiment takes place. In this case, very thin (that is, short exposure time) CuO films 104 may show appreciable Cu?0 structure due to a monolayer of Cu-,0 at the interface. In this regard, it should be re- emphasized that even a monolayer can produce significant changes in the differential reflectivity. In the case of CuO (see Figure 22) , a change in thickness of a single o monolayer (about 5 A) can cause a change in the dif- ferential reflectivity of several percent. Such a change is very readily detected with the present instrument. As a prelude to true in situ measurement of the dif- ferential reflectivity, certain experiments were under- taken to monitor changes in film properties with changing experimental conditions. Figure 26A shows a differential reflectogram of a copper sample potentiostated in a nil chloride solution of pH 10 at +500 mv SCE for sixteen hours. This potential is well into the CuO region of the Pourbaix diagram. The resulting spectra is similar to Figure 14, and is characteristic of CuO. After measure- ment, half of the sample which serves as a pure copper reference, was recoated with Miccroflex lacquer, and the sample immersed in fresh electrolyte. This time, how- ever, the sample was potentiostated in the Cu~,0 region of the Pourbaix diagram, at -225 mv SCE. After twenty hours the sample was removed and remeasured, producing the 105 106 spectra of Figure 26B. New structure, characteristic of Cu?0, is evident at 320 and 380 nm, superimposed on the strong CuO background. The clean copper side of the sample was again recoated, and the sample again potentio- stated at -225 mv for an additional sixty-eight hours. When the sample was measured for the third time, the curve of Figure 26C was produced, showing clear evidence of extensive Cuo0 film formation. The peak positions and even their relative intensities are characteristic of Cu20 The differential ref lectometer yields other informa- tion on the films formed in Figure 26 in addition to their identity. The zero lines of Figure 26 have been shifted for clarity, but a superposition of the three curves indicates that the CuO pattern remains relatively constant, while the Cu?0 pattern "grows" on top of it. Since the intensity of the CuO structure is directly pro- portional to the thickness of the film (see Figure 22) , it is reasonable to assume that the thickness of the CuO film does not change substantially when the potential is shifted to the Cu?0 area of the Pourbaix diagram. Hence, the shift in potential induces growth of a Cu20 layer under the existing CuO layer, rather than the dissolution of the CuO layer. This is significant observation which could 107 not be readily obtained with strictly electrochemical techniques . In Situ Experiments One of the major advantages of the differential re - flectometer compared to other techniques is its ability to perform in situ measurements. Theoretical considerations, explained in the experimental procedures section, indi- cate that an electrolyte surrounding the sample and the presence of a window to the corrosion cell should not introduce additional structure to the differential re- flectogram. Figure 27 shows two differential reflecto- grams of samples potentiostated in the Cu?0 region of the Pourbaix diagram (-200 mv SCE , pH 9.2) for five and one- half hours. One is the spectrum of a sample prepared in the normal manner, where Miccroflex covered half the specimen surface. The other represents a sample prepared in collodion covering half the specimen surface and analyzed in situ- The sensitivity, or amount of amplifi- cation, for the two curves is identical. Three points stand out. 1) The peak positions of each curve are at the same wavelengths. 2) There is no additional structure introduced in the in situ spectrum. 108 109 3) There is some loss of intensity or amplitude of the in situ curve compared to the non-in situ curve, and an associated increase in the noise level. The first two points are most important because they establish the fact that this technique of making measurements gives results which are consistent with simpler measurements not performed in situ. The third point indicates that absorption of the light as it passes through the window, electrolyte, and collodion, as well as losses due to reflection at the air/window and window/ electrolyte interface and small amounts of turbulence in the electrolyte, combine to produce a somewhat less intense and more noisy signal. This effect is not un- expected, nor is it troublesome as long as no additional structure is introduced in the differential ref lectogram. The spectra of Figure 27 were made after only five and one-half hours in order to minimize the degradation of the collodion coating on the in situ sample. In order to investigate the effect of the collodion on the inten- sity of the spectra, an additional experiment was per- formed. First, a sample was prepared in the usual manner in the Cu^O region of the Pourbaix diagram. The sample was removed from the electrolyte, dried, and the Miccro- flex stripped off. Figure 28A represents the differential 110 reflectogram made from this sample. The sample was placed in the dry in situ corrosion cell and another reflectogram made, Figure 28B. The lack of significant changes in the spectrum shows that the effect of the presence of the window is negligible. (The zero lines of Figures 28A and 28C have been shifted for clarity.) Next the in situ cell was filled with electrolyte and a third differential re- flectogram taken immediately, before the thin film cor- rosion product could be altered to any great extent. This spectrum is shown in Figure 28C, which shows how an in situ differential reflectogram would appear if there were no collodion present on the "pure" copper side of the specimen. Compared to Figures 28A and 28B, there is a general sharpening of the peaks and an increase in in- tensity. These changes can be attributed to the coupling effect of the electrolyte, which acts as a light pipe to optically couple the window and sample more efficiently than when the sample is surrounded by air. The index of reflection of the corrosion product and the quartz window is more nearly matched by the index of refraction of the electrolyte than that of air, resulting in a more ef- ficient optical coupling of the two surfaces. A similar effect is used in oil immersion optics. Here it partly compensates for the absorption of the light beam by the Ill c 60 c o nC c '/) ■H „ T3 s rt r— 1 rt 3 u > >. o +j u IH aj rH rt Q) x ,c r — N •M U 0) v ' r— 1 fi ft •H 0) s +-> rt ~ >- t/i , — > rH < o rt > — ' u e C rt ■H^ »H +-> bflT3 -H O H u 3TJ a; to a) i— i rt i— i m •^ 112 collodion. Without this coupling effect, uniform absorption of the light by the collodion would decrease the intensity of in situ measurements relative to non- in situ measurements more than is indicated by Figure 27. Several experiments have been performed to monitor the growth of corrosion films in situ. One example is shown in Figures 29 and 30 which represent the evolution with time of the differential ref lectograms of a sample potentiostated in the Cu-O region of the Pourbaix diagram at pH 9.2 and a potential of -200 mv SCE. These curves are typical for samples potentiostated under these condi- tions, although in this case the collodion covering the "clean" copper resisted breakdown by the electrolyte for an unusually long time. Immediately after immersion (t = 0) , the differential ref lectogram shows a reasonably flat spectrum, with no evidence of structure. The dis- placement from the zero line actually decreases during the first three hours after immersion, perhaps showing the dissolution of a previously formed oxide. It is also possible that the decrease in intensity is due to inter- ference effects discussed in the theoretical section. This was probably not the case, however, since no further oscillations were observed. In addition, interference o oscillations would imply a film thickness of over 500 A, 113 ES +-> 03 DO (D -H 03 +-> X (/) -H O 03 •H ,n +-> ^ G 3 CD O ■P Cl, O P-, CD B o 03 0 O c/) 4-> -H 0) ,c; o3 oj •H . +-> CTl G en ac 0) m w •H CJ T3 CO 3 > £ CXI Si 00 114 O 1^- -C IT) -C <*- CD if) CD o CD >H ro ■P P O CD •xi CTi 4-> o CXI o ,P P V U cU W i— 1 P O K O V C Is o O CD O 4H o X, T3 h- P n3 00 CD m 0) 4-> O > MH Si >H P P ,P O U oo •H P CD P o3 ,p 0) P P CD P & ■H CD P U O -H 03 UOOX 115 which is generally not observed for Cu?0 in aqueous solu- 20 tions. Finally, few in situ experiments have shown a similar decrease in intensity. This implies that the initial anomalous intensity decrease is caused by the dissolution of a previously formed film, which may occasionally be produced under conditions such as lengthy exposure to the air, high humidity, and so on. After about three hours, typical Cu?0 structure be- gins to appear. The amplitudes of the peak increase with time, and the curves more nearly approach those of "standard" Cu90 patterns, which were not measured in situ. The intensity of the mean peak near 385 nm as a function of time was fit to linear, power, logarithmic, and exponential functions by the method of least squares. The spectrum taken after three hours was used as a zero line because of the presumed dissolution of a previously formed film. The best fit was provided by a power curve — 70 of the form AR/R[|] = .62t which had a correlation ? coefficient of r = .99. If a theoretical curve of AR/R" versus thickness were available for Cu?0 as in the case of CuO (Figures 17 and 22) the kinetics of the growth of the oxide could be readily determined. Experiments to monitor the growth of CuO films were not successful due to problems with breakdown of the 116 collodion protective layer after short exposure times. The relatively high potentials applied to the sample when it is potentiostated in the CuO region of the Pourbaix diagram appear to accelerate the onset of crevice cor- rosion under the film. Further Discussion When using the differential ref lectometer in cor- rosion work, some way of matching the corrosion product and a specific differential reflectogram must be pro- vided. Perhaps the most elegant method is to calculate the differential reflectogram from the optical constants and thickness of the film. This method, described in detail earlier in this section, works well when the opti- cal constants of the film are known as a function of wavelength. If these data are not available, as in the case of Cu-,0, other methods must be employed to match the compounds and the spectra. One technique is to potentiostat a sample well into a specific region of the Pourbaix diagram and assume that the predicted species is the one which is observed. This technique probably works well in most cases, but is rather limited from a research point of view. Another is to grow a thick layer of some corrosion product, analyze it using X-ray diffraction techniques, 117 and correlate the results with a differential reflecto- gram of the film. Thinner films, which would have the same differential ref lectogram, could be identified even though they themselves were too thin for X-ray analysis. In order to test this idea, Cu70 and CuO films were prepared by potentiostating copper in the normal manner for periods as long as a week. X-ray analyses at the University of Florida failed to detect any trace of the corrosion product. The samples were sent to Dr. B. Cavin at Oak Ridge National Laboratories where a special X-ray apparatus employing monochromatic X-rays and digital signal averaging techniques was used in an attempt to detect the films. After runs as long as seventy-two hours, the only X-ray lines which were ob- served other than those belonging to copper were the strongest line of Cu20 and the strongest line of CuO. Both peaks were detected on both samples, and both peaks were so weak as to be barely distinguishable from back- ground. It is likely that the time lag between the forma- tion of the films and their measurement enabled both oxides to form on both samples. In any event, X-ray analysis does not seem to be a promising technique for confirming the identity of these films. One final technique which may be used to confirm the identity of a corrosion product is to match its band 118 structure or optical reflection spectrum with structure evident in the differential ref lectogram. This technique is useful in the case of Cu~0, because of the extensive optical spectroscopy and band structure calculations which have been performed on that particular • , 20, 30-32,35-47 oxide. ' ' The band diagram of Cu~0 which is shown in Figure 5 reveals a band gap of 1.77 eV. According to these cal- culations the smallest energy for optical interband transitions should therefore be 1.77 eV. It is generally realized that the accuracy of band calculations is between 0.5 and 1 eV and depends strongly on the method of calcu- lation and the input data. The experimental value for the 20 band gap was found to be 2.2 eV, which, in the light of the preceding statements, can be considered to be in fair agreement with the calculations. The differential reflectogram of Cu?0 shown in Fig- ure 13 shows structure at 2.25 eV which, as stated above, is identical to the energy associated with the band gap. Therefore, the structure at 2.25 eV can be tentatively attributed to the onset of interband transitions. From the calculated band diagram (Figure 5) one may deduce that the next set of interband transitions takes place approxi- mately 1 eV above the threshold energy. This agrees with 119 our observations, where the second peak in the dif- ferential reflectogram occurs at 3.26 eV (380 nm) . Following this transition, the calculated band dia- gram (Figure 5) predicts a multitude of possible closely spaced transitions at higher energies. This is confirmed by the differential reflectogram (Figure 13) which shows complex and broad structure in the 3.4 to 5.4 eV energy interval (365 nm to 2300 nm) . Experimental reflectivity data, shown earlier in Table 2, shows essentially the same picture. Table 5 is a reproduction of Table 2 with the results of this study added for comparison. The results evident from this table, in addition to the prediction of the Pourbiax diagram and the results of the X-ray analysis, seem to establish beyond reasonable doubt that differential re- flectograms typified by Figure 13 represent Cu-O- Another complicating factor which can cause errors in the analysis of thin film corrosion products when using optical techniques is the effect of surface roughening. The equations used in ellipsometry , or for that matter differential ref lectometry , assume specular reflectance from the metal substrate. Surface roughening causes scattering of the light beam and can interfere with ellipsometric measurements. Presumably the same sort of 120 o (J ( ( > 3 ■M LO r-l LO rH t"s r-~ I/) (/■; CN cni Tj- 00 to 1— 1 Tt to LO •rH u X CD X ^ Pi ■M +-> -M O 0) E e o CN o +-> to r< u •+H CD Tj [^ vO vO to LO CT, CD c-^ OO r i VO to LO rH +j 13 Mh rt i— ( (XI to to Tt «* CD CD rH C CC 3 •H O r— 1 LO T* T* I— 1 to !>. rt p-t CNl to r-- rH oo O ■m rt rt J3 -H u CNl to to T* «* LO O -P rt cd Jh r< o +-> CD <-> rx. u m 0 m 0\ IO \D vO LO O CNl vO CNl to i—( Pj-H 4-> lo r^ v£5 t~~ to Tt t^~ r^ en O CNl CO Q O^ ' "« £* CNl (X) to to Tfr T* Tl" Tt Tt LO LO o o r-i-M 3 CD 0-> ^o o^ LO 0O LO 00 LO •H 03 M tO \Q (-> CTl O CN] PL, CD S to to to to *fr Tl- Sh O 3 U +-> W (J to 3 CD in 3 4-> cro CO -H to r--. c--- Tt oo to PJ ■H LO ■rf- r-~ to r-- o u CNl tO to T}" Tt LO CD i— 1 l+H CD CC 121 problems should exist with differential ref lectometry . Surface roughening due to incomplete polishing causes no problem in differential ref lectometry because both halves of the sample undergo the same polishing procedure, so any effect due to surface roughening should cancel out. However, roughening caused by electrochemical means acts only on the unprotected half of the specimen, and hence should contribute to the differential ref lectogram. Since no peak shifts or additional structure were observed when samples were potentiostated for extended periods of time, however, it may be reasonably concluded that any surface roughening effects which did occur were insuf- ficient to cause any significant changes in the dif- ferential reflectivity. Recently a certain problem has arisen with regard to a few samples prepared potentiostatically when Miccroflex is used to protect the reference half of the sample. The differential ref lectograms which are taken from these samples show Cu?0 structure, but with inverted polarity. This implies that the "protected" side of the sample is corroding preferentially and protecting the exposed side. The Cu^O spectra are observed regardless of whether the sample is potentiostated in the Cu?0 or CuO area of the Pourbaix diagram. The reason appears to be a lack of adherence to 122 the copper by the Miccroflex, resulting in a concentration cell in the gap between the Miccroflex and copper. This increases the acidity of the electrolyte trapped under the Miccroflex and drives the covered area anodic with respect to the rest of the sample. Both effects tend to move the area covered with Miccroflex into the Cu?0 area of the Pourbaix diagram. This results in a film of Cu20 being formed on the supposedly "clean" half of the sample, while the exposed half of the sample is relatively pro- tected. The solution to the problem is simply to obtain fresh lacquer. Finally, it was mentioned in Section II that elec- troref lectance effects can complicate thin film corrosion studies which employ modulation spectroscopy. Effects due to modulation of the electrode potential, such as are observed in electrochemical modulation spectroscopy (EMS), are of no concern here since the sample potential is not modulated. Effects due to the electric field emanating from a sample which has a DC charge on it have 22 ? 7 also been observed. ' This field alters the optical constants of the electrolyte near the sample surface, and could affect measurements made in situ. However, the electric field surrounding the samples used in these experiments must be fairly uniform, particularly near the 123 center of the flat face which is being observed. Here again the differential nature of the technique would tend to cancel out any modifications to the index of refrac- tion of the electrolyte, which must be small in any event. The fact that there is no additional structure in Figure 27B (measured in situ) compared to Figure 2 7A (not measured in situ) seems to confirm this reasoning. SECTION V CONCLUSIONS AND SUGGESTIONS FOR FURTHER WORK It is hoped that this work shows that the application of differential ref lectometry to the study of corrosion phemonema, even at its present early stage of development, can serve as a valuable supplement to the more established techniques of corrosion science. It has been shown that differential ref lectometry can verify the existence and identity of thin film corrosion products on metal sub- strates. Different corrosion products have been shown to produce unique differential ref lectograms , which can be qualitatively predicted from reflectivity data and band diagrams, or quantitatively predicted from the optical constants of the film. As in the case of ellipsometry , the thickness of the film may be determined if the optical constants are known. The technique is most useful in the o film thickness range 0 to 200 A, which covers the range of most passive and semipassive films. This range is often too thick for most surface analysis techniques which do not require a vacuum, and too thin for X-ray analysis. Finally, it has been shown that differential reflec- tometry is capable of in situ measurements of thin film 124 125 corrosion products. More importantly, the electrolyte does not add additional structure to complicate the analysis, so measurements in optically dense electrolytes are pos- sible. There are many interesting possibilities for future research. For example, the well-established techniques of compositional modulation " can be used to investi- gate dealloying, especially the dezincif ication of brass. Small compositional changes, which are difficult to mea- sure using conventional techniques, could be easily monitored in real time while the sample was still in the electrolyte. In order to facilitate such in situ studies, additional work needs to be performed on the development of transparent, protective coatings for the reference half of the sample. They must be able to be applied thinly and evenly, and must have uniformly low absorption throughout the range of the instrument as well as being immune to attack by the electrolyte. More calibration curves of film thickness versus differential reflectivity need to be calculated. The curves for CuO shown in Figures 17 and 22 were calculated for a wavelength of 600 nm, but any wavelength within the range of the instrument could be used. In retrospect a wavelength of approximately 400 nm would be better for 126 the case of CuO, since the differential reflectivity is higher at that wavelength than at 600 nm for all thick- nesses in the range of interest. As a consequence, greater precision could be obtained in the determination of the thickness of a film. Similarly, the calibration curve should be made near the strong peak at 385 nm in the case of Cu20, when the optical constants of that com- pound are determined. Wavelengths shorter than approxi- mately 300 nm should be avoided, since the output of the Xenon light source is not strong in that area. Additional metal/electrolyte systems could be investigated using differential ref lectometry , with the aim of developing experimental Pourbaix diagrams for commercial alloys whose theoretical Pourbaix diagrams are not easily calculated. Here the technique can be a valuable supplement to electrochemical techniques, especially when the corrosion product cannot be exposed to vacuum. REFERENCES 1. M. Fontana and N. Green, Corrosion Engineering (McGraw-Hill, New York, 196 7) 2^ 2. E. D. Verink, Jr., personal communication. 5. B. D. Cahan and R. F. Spanier, Surface Sci . 16 (1969) 166. 4. J. A. Holbrook and R. E. Hummel, Rev. of Sci. Instr. 4-4 (1973) 463. 5. E. D. Verink, Jr., and M. Pourbaix, Corrosion 27 (1971) 495. 6. P. C. Ladelfe, A. W. Czanderna, and J. R. Biegen, Thin Solid Films 1_0 (1972) 403. 7. 0. S. Heavens, Optical Properties of Thin Solid Films (Dover, New York, 1965). 8. W. K. Paik and J. O'M. Bockris , Surface Sci. 28 (1971) 61. 9. R. C. O'Handley, Surface Sci. 4_6 (1974) 24. 10. R. C. O'Handley and D. K. Burge , Surface Sci. 48 (1975) 214. ' 11. S. Gottesfeld and B. Reichman, Surface Sci. 4_4 (1974) 377. 12. Gaertner Scientific Co., Chicago, 111., Bulletin EH. 13. Gaertner Scientific Co., Chicago, 111., Bulletin EA203-76P. 14. S. N. Jasperson, D. K. Burge, and R. C. O'Handley, Surface Sci. 37_ (1973) 548. 15. H. J. Mathieu, D. E. McClure , and R. H. Muller, Rev. Sci. Instr. 45 (1974) 798. 127 128 16. R. M. A. Azzam, Rev. Sci . Instr. 4J7_ (1976) 624. 17. J. D. E. Mclntyre and D. M. Kolb , Symposium of the Faraday Society 4 (1970) 99. 18. J. D. E. Mclntyre and D. E. Aspnes, Surface Sci. 24 (1971) 417. ' ' — 19. D. M. Kolb and J. D. E. Mclntyre, Surface Sci. 28 (1971) 321. ' ' — 20. J. Kruger, J. Electrochem. Soc. 10_8 (1961) 503. 21. S. Matsuda, K. Sugimoto, and Y. Sawada, Trans. Japan Inst. Metals 1_8 (1977) 66. 22. J. D. E. Mclntyre, Surface Sci. 37_ (1973) 658. 23. J. D. E. Mclntyre and W. F. Peck, Jr., in: Advances in Electrochemistry and Electrochemical Engineering , Vol. 9, Ed. R. H. Muller (Wiley- Interscience , New York, 1973) 66. 24. J. D. E. Mclntyre, in: Optical Properties of Solids - New Developments, Ed. B. 0. Seraphin (North Holland, New York, 1976) 555. 25. B. D. Cahan, J. Horkans, and E. Yeager, Surface Sci. 3_7 (1973) 559. 26. J. Horkans, B. D. Cahan, and E. Yeager, Surface Sci. 46 (1974) 1. 27. J. D. E. Mclntyre and W. F. Peck, Jr., Faraday Disc. Chem. Soc. 5_6 (1973) 122. 28. W. R. Hansen, in: Progress in Nuclear Energy, Series IX, Anal. Chem.; Vol. 11, Emittance and Reflectance Spectroscopy, Eds. H. A. Elion and D. C. Stewart (Pergamon, New York) 1. 29. J. D. E. Mclntyre and W. F. Peck, Jr., in: Proc. Symp. on Electrocatalysts , San Francisco, 1974~ Ed. M. W. Breiter (The Electrochemical Society, Princeton) 212. 30. A. G. Zhilich and V. P. Makarov, Soviet Phys . Solid St. 3 (1961) 429. 129 31. R. J. Elliott, Phys. Rev. 124. (1961) 340. 32. J. P. Dahl and A. C. Switendick, J. Phys. Chem. Solids 2^ (1966) 931. 33. W. Kohn and N. Rostoker, Phys. Rev. 9_4 (1954) 1111. 34. J. C. Slater, Phys. Rev. 5JL (1937) 846. 35. S. Brahms and S. Nikitine, Solid State Comm. 3 (1965) 209. 36. M. Balkasanski, Y. Petroff, and D. Trivich, Solid State Comm. 5_ (1967) 85. 37. B. Prevot, C. Carabatos, and M. Seiskind, Phys. Stat. Sol. 10 (1972) 455. 38. M. Zouaghi, Phys. Stat. Sol. 11 (1972) 219. 39. M. Zouaghi, B. Prevot, C. Carabatos, and M. Seiskind, Phys. Stat. Sol. 11 (1972) 449. 40. S. N. Shestatskii and V. V. Sobolev, Phys. Stat. Sol. 32_ (1969) K109. 41. S. N. Shestatskii, V. V. Sobolev, and N. P. Likhobabin, Phys. Stat. Sol. 42_ (1970) 669. 42. A. Daunois, J. L. Deiss, and S. Nikitine, Solid State Comm. 10 (1972) 649. 43. Y. Ebisuzaki, J. Applied Phys. 32_ (1961) 2027. 44. H. Wieder and A. W. Czanderna, J. Phys. Chem. 66 (1962) 816. ' — 45. J. Kruger, J. Electrochem. Soc. 1_0_6 (1959) 847. 46. J. Kruger, J. Electrochem. Soc. Ill (1964) 1038. 47. D. J. G. Ives and A. E. Rawson, J. Electrochem. Soc. 1_09 (1962) 458. 48. R. E. Hummel, J. A. Holbrook, and J. B. Andrews, Surface Sci. 37 (1973) 717. 130 49. R. E. Hummel and J. B. Andrews, Phys. Rev. B 8_ (1973) 2449. 50. R. J. Nas tasi -Andrews and R. E. Hummel, Phys. Rev. ] in press . 51. R. Enderlein, R. E. Hummel, J. B. Andrews, R. J. Nastas i-Andrews , and C. W. Shanley, to be published . 52. M. Pourbaix, Atlas of Electrochemical Equilibria in Aqueous Solutions, First English Edition (Pergamon Press, Oxford, 1966). 53. Thaddeus S. Lee, Master's Thesis, University of Florida, 1972. 54. L. G. Schulz, J. Op. Soc . American, 44_ (1954) 357. 55. L. G. Schulz and F. R. Tangherlini , J. Op. Soc. America 44_ (1954) 362. 56. D. Halliday and R. Resnick, Physics (Wiley, New York, 1968) 1086. 57. M. Hansen, Constitution of the Binary Alloys (McGraw-Hill, New York, 1958)! BIOGRAPHICAL SKETCH Charles William Shanley was born March 20, 1949, in Chicago, Illinois. He attended Loyola Academy in Wilmette, Illinois, where he graduated in 1967. Beginning in 1967, the author studied physics at the University of Dayton, Dayton, Ohio, where he graduated with a Bachelor of Science degree in 1971. He entered the graduate school at the University of Dayton in the field of materials science in 1971, and graduated in 1973 with a Master of Engineering degree. The author entered graduate school at the University of Florida in June, 1973, in the field of materials science and engineering. He received the degree of Master of Science in March, 1975, and has pursued the degree of Doctor of Philosophy since that date . The author is a member of the American Institute of Mining, Metallurgical, and Petroleum Engineers, the American Society for Metals, and Alpha Sigma Mu. 131 I certify that I have read this study and that in my opinion it conforms to acceptable standards of scholarly presentation and is fully adequate, in scope and quality, as a dissertation for the degree of Doctor of Philosophy. Hummel , Chairman Professor of Materials Science and Engineering I certify that I have read this study and that in my opinion it conforms to acceptable standards of scholarly presentation and is fully adequate, in scope and quality, as a dissertation for the degree of Doctor of Philosophy. 3?>k J. J. Hrdn Professor of Materials Science and Engineering I certify that I have read this study and that in my opinion it conforms to acceptable standards of scholarly presentation and is fully adequate, in scope and quality, as a dissertation for the degree of Doctor of Philosophy. E. D. Verink, Jr. Professor of Materials Science and Engineering I certify that I have read this study and that in my opinion it conforms to acceptable standards of scholarly presentation and is fully adequate, in scope and quality, as a dissertation for the degree of Doctor of Philosophy. :pinKKy Professor of R This dissertation was submitted to the Graduate Faculty of the College of Engineering and to the Graduate Council, and was accepted as partial fulfillment of the require- ments for the degree of Doctor of Philosophy. December 1977 Dean , College of Engineering Dean , Graduate School UNIVERSITY OF FLORIDA 3 1262 08553 1126