2 ,. .31! ”cf. r . . .1‘ D) a q .5 1 5553's.}: . 5 a .prno.5...n. _ . . t... . all} 5.3:. .2. 5‘ fig. 1.! . 17.3 I.“ 'u' v1, ut! "3; 3‘3“. 4.5.1.. 3:. A» .v. 51...”: 393.}...5); .7 a .11.. 11341:! 1).). ..:»I.... (If .. ‘5' meUc‘ ' l ‘.I‘, g, - - C l . \- ..l’ I!» 1):....)II:IO-.l;oil§.w.' ..:L:..... 2.2:. fix. :3. 1.4.. » A€¢al .nlt‘,> “ .‘I? O t) Oa'av'J’I-v A, .l}(J-v'. I‘VL- i‘vl .- u L . . , 2" l‘ 4. .01 [I .u, bio. 15,1. :y)..%‘ 9.: . . , , ’1... ~51}... x... n I; til .1}? ‘ incl. «.Vv13t . . LSJK t silvv. » . (I. J . . I.) Illlfneui‘ '- . . :2 up: . , . fiddhyhfivfunfindbr. 2. .11;5..!I.v5aunw I . . . ahi . 4‘er . ., . ‘ . :1}... .. . . . . ‘ . , 1.5» 31!: r I. .zu..?....;..1». #2.!A1hufihflwwmk.,u...mwm.ndmmur,. ,1 5. - 71......ng 1...“... , . 1 t ‘9 ’ it. .. . ‘ .. . Fl i u .7 1 .tl‘ V . 2 , . Jew: EV... .. at: 41V: swummauMcvmwfi~ z . . .I I4 12.x I 14"». ‘ID I, MICHIGAN STATE UNIVERSITY LIBRARIES L- 1%qbseeb 11 11111 11111111111111111111111111 t uamv 11 1 ”I“: m" SW 3 1293 00592 1295 Thisistocertifythatthc dissertation entitled I . ABSORPTION- CORRECTED FLUORESCENCE THROUGH FIBER OPTICS II. DESIGN AND ANALYSIS OF NONRECURSIVE DIGITAL FILTERS presented by Thomas Patrick Doherty has been accepted towards fulfillment of the requirements for Ph . D . Chemis try ‘degreein afi?fi%;4. Major professor Datew MS U is an Afl‘innatiw Action/Equal Opportunity Institution PLACE IN RETURN BOX to remove thls checkout from your record. TO AVOID FINES return on or More due duo. P_________________———-————-— DATE DUE DATE DUE DATE DUE I I I I I _, MSU Is An Allirmdive Action/Emmi Opportunity Inditution # 7 _—_—___ I. ABSORPTION-CORRECTED FLUORESCENCE THROUGH FIBER OPTICS II. DESIGN AND ANALYSIS OF NONRECURSIVE DIGITAL FILTERS By Thomas Patrick Doherty A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1990 9001944 ABSTRACT I. ABSORPTION-CORRECTED FLUORESCENCE THROUGH FIBER OPTICS II. DESIGN AND ANALYSIS OF NONRECURSIVE DIGITAL FILTERS m Thomas Patrick Doherty I Several instrumental and mathematical methods have been developed that correct fluorescence measurements for the inner filter effect. Almost all are designed for use with 1 cm cuvettes, and are thus incompatible with flowing sample streams. A fiber optic fluorometer was constructed that addresses this problem. The instrument simultaneously measures the fluorescence at the front and rear of a cylindrical flowcell, as well as the transmittance of the excitation beam. From the absorbance at the excitation wavelength and the ratio of the fluorescence at the front of the cell to that at the rear, the absorbance at the emission wavelength is estimated. The computer that controls the instrument uses these two absorbances to calculate what the fluorescence measurement would be in the absence of inner filter effects. The corrections are useful to a total absorbance of 2.5. The equation that is used to calculate the corrected fluorescence contains an optical transfer function that describes the propagation of light through a fiber optic bundle. This transfer function was originally derived empirically, based on the measured fluorescence of a series of solutions of known absorbance. An expression is derived from first principles that matches the shape of the empirical function under certain conditions. The expression could prove useful for modifying the correction equation should the refractive index of the sample, or the wavelength of excitation or emission be changed. II The Savitzky-Golay filter familiar to many chemists is one example of a nonrecursive digital filter. Techniques for examining the frequency response of a nonrecursive filter are described, and are used to study three types of filters, including sliding window filters that can be used at the extremes of a data set. These are shown to have undesirable gain and phase properties that could lead to the erroneous interpretation of data, especially if they are applied repeatedly. Techniques for the design of a filter with a given frequency response are also described. Two types of filters are designed: a bandpass filter that removes correlated noise, and IOWpaSS filters for use when the sampling rate of a digitized signal is changed. To Paul and Joan, who know I can and to Linda, who is my inspiration ACKNOWLEDGMENTS None of the work described in this dissertation would have been possible without the support and guidance of my research advisor, Stan Crouch. I am perhaps most indebted to him for something which is not embodied in this text: the knowledge I gained because I was given the freedom to pursue any interesting idea, whether related to the projects described here or not. It is this "education” that made coming to Michigan State University truly worthwhile to me. That, and some memorable Lions games. I would also like to thank the other members of my committee: James Dye, Jack Holland, and especially my second reader, Chris Enke. Every time I spoke with him I was filled with a renewed sense of excitement at the many ways we can use computer software to get more information from our spectroscopic data. My list of mentors would not be complete without those who helped me decide to go to graduate school: John Zimmerman, Professor of Chemistry at Wabash College, and Garry Buettner, formerly at Wabash. Due to the size of the Crouch group during my stay, it is not possible to thank all the members of my research group who helped me at State. But you know who you are. I thank all of you for advice and help in the lab, and friendship inside it and out. However, several members of the research group must be mentioned separately because their work directly affected what is described herein. Pete Wentzell was my chief collaborator in the digital filter work. Without his focus and clarity it would have remained merely a hobby and never would have made it into this dissertation. I am also gateful to Cheryl Stults and Eric Erickson for providing other interesting applications to try out my filters on. Pete’s counterpart on the fluorescence side was Mark Victor. We argued incessantly over the finer points of the model presented in Chapter 2, which ultimately led to geat improvements. Mark also provided plenty of experimental help: be refined the design of the stepper motor controller and built the final version, and designed the fiber Optic interface system for the PTR monochromators. My thanks also go out to Deak Watters and Russ Geyer in the machine shop who built the components of the interface. Max Hineman served as my main opties, math and quantum chemistry resource, and was also a geat roommate for three years. The members of the lunch bunch: Pete, Max, Dean Peterson, Kim Ratanathanawongs, and Bob Harfmann, provided something much more valuable and lasting than research papers: memories. Ditto for all those with whom I picnicked, camped, skied, and played softball and volleyball. For reasons not beyond my control, I am able to thank the expediter of this dissertation, Larry Bowman, with whom I also had many interesting discussions about fluorescence. I owe an immeasurable debt to my parents, Joan and Paul. They always knew I could do it, even when I didn’t believe it. Finally, to the person to whom I owe the most, my wife Linda: thank you. Ifyou hadn’t come to Michigan State, I’d still be there. vi TABLE OF CONTENTS List of Figures _ X: List of Tables -xvi I. Absorption-Corrected Fluorescence Through Fiber Optics Chapter 1 : Introduction to Absorption-Corrected Fluorescence ...................................... 1 Review _ - ........................ 4 Mathematical Corrections .......... ..... 4 Instrumental Corrections .................... 5 Applications _ - - - ................ 6 Miscellaneous Phenomena Related to the Inner Filter Effect ................ 7 Absorption Correction For Flowing Sample Streams - -- - ........... 8 References----- - ............................................... 10 Chapter 2 : A Fluorescence Intensity Expression - 12 Review .......... 12 Fiber Optics ................................... 13 A Fluorescence Intensity Expression - - -- - - ............. 15 Summary of Previous Work - - - -- 17 Comparison of Theoretin and Experimental Models _ .......... 18 Conclusions --22 References ......... 24 Chapter 3 : Experiments in Absorption-Corrected Fluorescence - .......... 25 Theory .......................................... 25 Experimental Details .................................... 27 Apparatus 27 The Effect of Primary and Secondary Absorbance on the Fluorescence Ratio ...... 31 Introduction - 31 Experimental Details 33 Reagents - 33 Procedure 33 Results and Discussion 35 A Test of the Correction Procedure - ................. 50 Introduction- - -_ ............................. 50 Experimental Details - . ................. 52 Reagents .................... 52 Procedure -- .... 52 Results and Discussion - ......... 52 Real-Time Performance 56 References 64 Chapter 4 : Perspectives, Conclusions, and Suggestions for Further Study .................... 66 Limitations of the Current Instrument and Suggestions for Improvement ....66 Comparison of the Front Surface / Rear Surface Fluorometer and Cell Shift Instruments - -72 Suggestions for Further Study -73 Modifications of the Current Instrument 73 Other Experiments in Absorption-Corrected Fluorescence .................. 76 References 82 viii 11. Design and Analysis of Nonrecursive Digital Filters Chapter 5: Introduction to Digital Filters -- 84 Analog Filters 84 Digital Filters 86 Linearity and Time-Invariance 87 Advantages of Digital Filters 90 A Brief History of Convolution Filters in Analytical Chemisry 92 Savitzky-Golay Filters 92 Other Symmetric Filters 95 Asymmetric Filters 96 Conclusions 96 References 97 Chapter 6 : Frequency Responses of Digital Filters 99 Examples 103 Savitzky-Golay Filters - 103 Digital Up/ Down Integration 104 Sliding Window Filters 112 Performance of the Sliding Window Filters 1 18 Conclusions 125 References -- 126 Chapter 7 : Design of Digital Filters 127 Theory 127 Examples 128 Low-Pass Filters 128 Differentiators 132 Why Design Filters? -136 Examples Correlated Noise Removal Decimation Interpolation Conclusions References 137 137 141 145 151 153 Figure 2-1. Figure 2-2. Figure 2-3. Figure 2-4. Figure 3-1. Figure 3-2. Figure 3-3. Figure 3-4. Figure 3-5. Figure 3-6. Figure 3-7. LIST OF FIGURES Optical fiber acceptance cone Overlap region of light emitted from and collected by two adjacent optical fibers: (a) schematic diagram, (b) head-on View .. Comparison of experimental and theoretical transfer functions. ............... 16 .............. 19 Effect of changes in the parameters used in the theoretical model: (a) change in numerical aperture (1 - 4 = numerical aperature of .1, .05, .025, .0125), (b) change in fiber radius (1 - 4 = fiber radius of .1, .3, .9, 2.7 mm). -- Cylindrical cell model. - .................................................. Diagram of fluorescence spectrometer. ............ Diagram of sample cell: (a) cutaway view, (b) view showing fiber optic windows. ............... 23 ............... 26 ............... 28 ............... 30 Output of the four detectors for 500 11M quinine sulfate (08), 0 to 910 11M bromocresol purple (BP): (3) reference channel, (b) transmittance channel, (c) front surface fluorescence channel, (d) rear surface fluorescence channel. .......................... Fluorescence ratio versus absorbance at 438 nm. 1= 250 um quinine sulfate (OS), 2= 500 uM 08, 3= 750 14M 08, 4= 1000 11M OS, 5- 1250 MM 08, 6-- 1500 uM QS. Parameter 0 versus absorbance at 353 nm. Parameter b versus absorbance at 353 nm. - ...... ............... 34 ............... 36 ............... 38 Figure 3-8. Figure 3-9. Figure 3-10. Figure 3-11. Figure 3-12. Figure 3- 13. Figure 3-14. Figure 3-15. Figure 3-16. Figure 3-17. Absorbance at 438 nm (1 - calculated, 2 - measured) versus sodium fluorescein (SF) concentration: (a) 250 um quinine sulfate (OS), (b) 500 um OS Absorbance at 438 nm (1 a calculated, 2 = measured) versus sodium fluorescein (SF) concentration: (a) 750 11M quinine sulfate (OS), (b) 1000 um OS Absorbance at 438 nm (1 = calculated, 2 = measured) versus sodium fluorescein (SF) concentration: (a) 1250 11M quinine sulfate (OS), (b) 1500 I‘M OS Front surface fluorescence intensity ( 1 = measured, 2 = corrected using calculated absorbance at 438 nm) versus sodium fluorescein concentration: (a) 250 um quinine sulfate (OS), (b) 500 um OS Front surface fluorescence intensity (1 = measured, 2 = corrected using calculated absorbance at 438 nm) versus sodium fluorescein concentration: (a) 750 um quinine sulfate (OS), (b) 1000 nM OS Front surface fluorescence intensity (1 8 measured, 2 a corrected using calculated absorbance at 438 nm) versus sodium fluorescein concentration: (a) 1250 11M quinine sulfate (OS), (b) 1500 14M QS Absorption spectrum of 0.909 11M bromocresol purple (BP) in 0.1 E HClO4. Absorbance at 438 nm (1 = calculated, 2 = measured) versus bromocresol purple (BP) concentration: (a) 250 11M quinine sulfate (OS), (b) 500 um OS Absorbance at 438 nm (1 = calculated, 2 - measured) versus bromocresol purple (BP) concentration: (a) 750 um quinine sulfate (OS), (b) 1000 11M OS 41 ---42 43 46 47 48 51 53 - ,_54 Fluorescence intensity (1 :- uncorrected, 2 = corrected using calculated absorbance at 438 nm, 3 - corrected using measured absorbance at 438 nm) versus bromocresol purple (BP) concentration: (a) 250 11M quinine sulfate (OS), (b) 500 14M 05 --57 Figure 3-18. Figure 3-19. Figure 3-20. Figure 4-1. Figure 4-2. Figure 4-3. Figure 4-4. Figure 5-1. Figure 6-1. Figure 6-2. Figure 6-3. Figure 6-4. Fluorescence intensity (1 = uncorrected, 2 - corrected using calculated absorbance at 438 nm, 3 - corrected using measured absorbance at 438 nm) versus bromocresol purple (BP) concentration: (a) 750 11M quinine sulfate (OS), (b) 1000 11M OS. filters in place. passes. 58 Real-time results: (a) measured primary absorbance, (b) estimated secondary absorbance 60 Real-time results: (a) front surface fluorescence, (b) corrected front surface fluorescence. - 61 Time profile of rear surface fluorescence measured while sample cell is rinsed with 0.1 E HClO4: (a) 0 to 4 minutes, (b) 0 to 25 minutes ...... -...68 Front surface fluorescence blank: (3) filters removed, (b) ............. 70 Calculated ratio of front surface fluorescence measurement from 2 cm flow cell to front surface fluorescence measurement from 0.5 cm flow cell. ............. 78 Flowing bubble system: (a) diagram showing time-varying pathlength, (b) calculated front surface fluorescence signal (1 a total absorbance negligible, 2 8 total absorbance of 2.0). 80 Illustration of the Nyquist theorem. ............ 88 Frequency responses of 11 point Savitzky-Golay filters: (3) linear, (b) quadratic, (c) quartic 105 Frequency responses of quadratic Savitzky-Golay filters: (a) 7 point, (b) 9 point, (c) 11 point. 106 Frequency responses of multiple passes of a quadratic 11 point Savitzky-Golay filter: (a) 1 pass, (b) 2 passes, (c) 3 - 107 Frequency response of 11 point quadratic first derivative Savitzky-Golay filter. 108 - Figure 6-5. Figure 6-6. Figure 6-7. Figure 6-8. Figure 6-9. Figure 6-10. Figure 6-11. Figure 6- 12. Figure 7-1. Figure 7-2. Figure 7-3. Figure 7-4. Figure 7-5. Figure 7-6. Normalized frequency responses of up/ down integration with rc=o.1. filtered filters: (a) 11 point, (b) 21 point, (c) 101 point. ....... 111 Frequency responses of quadratic initial point filters: (a) 9 point, (b) 11 point, (c) 13 point. - 114 Phase response of quadratic initial point filters: (a) 9 point, (b) 11 point, c) 13 point. 116 Frequency response of 11 point initial point filters: (a) linear, (b) quadratic, (c) cubic. 117 Frequency response of quadratic 11 point filters: (a) minus first point, (b) initial point, (c) third point. -119 Effect of sliding window filter on noise-free sine wave. .. 121 Effect of sliding window filter on noise - _ ....... 123 Effect of initial slope filter on st0pped-flow data: (a) stopped-flow transient, (b) calculated derivative. 124 Frequency responses of unwindowed 11 point filters: (a) fc = 0.1 times sampling frequency, (b) fc = 0.2 times sampling frequency, (c) fc= 0.3 times sampling frequency. 131 Frequency responses of the windowed versions of the filters shown in Figure 7-1. ............ 133 Frequency response of windowed 21 point derivative filter 135 Peak obtained on flow injection analyzer. (a) unfiltered, (b) 138 Frequency responses of (a) low-pass filter, (b) high-pass filter, (c) notch filter 140 Decimation by a factor of two. ............ 142 xiv Figure 7-7. Figure 7-8. Figure 7-9. Figure 7-10. Figure 7-11. Time-of-flight mass spectrum of xenon: (a) original spectrum, (b) decimated by 10, no filtering. (3) Frequency response of decimation filter. (b) Filtered time-of-flight mass spectrum of xenon; filtered, then decimated by 10 Interpolation by a factor of two Interpolated Gaussian profile: (a) before filtering, (b) results of Savitzky-Golay interpolation filter, (c) results of windowed ideal low-pass interpolation filter. Frequency responses of decimation filters: (a) windowed ideal low-pass filter, fc - 0.477, (b) 21 point eighth order Savitzky-Golay filter. 144 146 148 - 149 150 Table 3-1. Table 3-2. Table 3-3. Table 3-4. Table 3-5. Table 3-6. LIST OF TABLES Comparison of measured and calculated secondary absorbance values absorbance 44 Comparison of linear regression results for measured and calculated secondary absorbance 44 Comparison of uncorrected and corrected fluorescence 49 Comparison of measured and calculated secondary 55 Comparison of linear regression results for measured and calculated secondary absorbance 55 Comparison of uncorrected and corrected fluorescence ...... 59 values CHAPTER 1 : INTRODUCTION TO ABSORPTION-CORRECTED FLUORESCENCE Fluorescence techniques are used in quantitative analysis when their sensitivity, selectivity, and wide dynamic range are justified. A fluorescence assay is typically 10-1000 times more sensitive than an absorption assay. Selectivity is provided in part by the ability of the analyst to select the excitation and emission wavelengths and in part by the fact that many compounds do not fluoresce with high efi'iciency (1). Measurement of the lifetime of a fluorophore or the polarization of its emission can also be used to enhance selectivity. In most quantitative applications it is assumed that the observed fluorescence intensity is proportional to the concentration of the fluorophore. In fact, the fluorescence intensity expression involves many terms associated with the fluorophore, the characteristics of the sample matrix and the geometry of the spectrometer (2). Although the geometry of the spectrometer is usually constant, changes in the nature of the sample from one measurement to the next can cause this assumption of linearity to be incorrect. The intensity of the total fluorescence emitted is assumed to be given by: FocIo(1 - 10 *1”) where F is the fluorescence intensity, I0 is the incident monochromatic radiant power, 6 is the absorptivity of the fluorophore, c is the concentration, and b is the pathlength, so that she - A, the absorbance of the fluorophore. In the limit of low absorbance this equation can be approximated by Forloebc 2 which is the basis of the assumption of linearity between the observed intensity and concentration, although it should be stressed that the expression does not apply to a typical measurement made in a spectrometer. The simplest example in which the obsen/ed fluorescence is not proportional to the fluorophore concentration occurs when the fluorophore absorbs significant amounts of radiation. This is often what limits the upper end of the linear dynamic range of a fluorescence calibration curve. Samples containing such high concentrations of fluorophore are typically diluted until the concentration lies within the linear range of the calibration curve (3). This is not a general approach for several reasons. In high throughput, automatic assays, either all of the samples must be diluted, possibly causing the concentrations of certain samples to fall below the detection limit, or none are diluted, in which case the concentration of other samples may be erroneously estimated. Also, dilution may upset the position of an equilibrium, or it may not be possible, as in the case of in situ measurements. In other cases the absorbance may be high at the excitation wavelength (primary absorbance) or the emission wavelength (secondary absorbance) due to other components of the sample matrix. This phenomenon is known as the 111' per filter effect. If this occurs, estimates of the concentration of a fluorophore may be in error if the matrix of the unknown sample absorbs a different amount of radiation than the matrix of the sample used to develop the calibration curve. In this case, a calibration curve is often generated using the sample itself via standard additions, but once again this is not a general solution. Standard additions are usually not possible with high throughput automated assays or in situ measurements. Furthermore, if the fluorophore is contributing to the inner filter effect, standard additions will not produce a linear calibration curve. 3 In other cases that fluorescence detection is used, the concentration of the fluorophore remains constant, and some species is added to the solution to change the emission intensity by some other means, such as a change in the equilibrium between fluorescent and non-fluorescent forms of the fluorophore, or a change in the quantum efiiciency due to quenching. The former technique can be used to determine the binding of substrates to enzymes. The latter technique is particularly popular for biochemical applications, since the extent of quenching is a function of polarity of the environment of the fluorescent site, the accessibility of the fluorophore to the quencher, and the proximity of acceptors of energy transfer (4). There are also some assays, particularly immunoassays, that are based on fluorescence quenching. In both cases, the change in fluorescence intensity is assumed to be due to the phenomenon in question and not to changes in primary or secondary absorbance. Reaction-rate methods sometimes employ fluorescence detection of a reactant or product. If the primary or secondary absorbance of the solution should change with time, or the reaction should start or end in the non-linear region of the calibration curve for the fluorophore, the estimate of the concentration of the fluorophore, and hence the reaction rate, could be in error. One example is the measurement of reduced nicotinamide adenine dinucleotide (NADH) in enzyme immunoassays (5). In some cases inner filter effects are unavoidable. For this reason, many researchers have proposed mathematical corrections of observed fluorescence values that require measurement of the absorbance of the sample at the excitation and emission wavelengths. Others have constructed spectrofluorometers that automatically produce absorption-corrected fluorescence spectra. The discussion above indicates that there is clearly a need for ways to make absorption-corrected fluorescence measurements. 4 REVIEW There is a long history of work in absorption-corrected fluorescence at Michigan State University, including the work of Holland et at (6-8), Christmann et at. (9-12), Adamsons et at. (13-16), and Ratzlaff et al. (17,18). Several of these researchers have written thorough reviews of absorption correction, and another good review has recently appeared (19). It is the purpose of this section to update those reviews. Mathematical Corrections Wiechelman, in addition to reviewing the field, has proposed an empirical mathematical correction for absorption (19). The difference between the fluorescence in the presence and absence of absorbers was fit by an equation of the form: AF = a(1 - e‘bA) where A is the sum of the primary and secondary absorbances measured on a spectrophotometer. Values of a and b were determined for several absorbers of biochemical interest added to several different proteins containing the fluorescent tryptophan residue. The parameters a and b were found to depend on the absorbing wecies, but not on the proteins. This study was interesting because the difference in the two fluorescence values, rather than the ratio, was fit by a curve. Batke has derived an expression for correction that can be used with fluorometers of side excitation, bottom emission design (20). The equation has been used to study the concentration dependence of the quantum efi'iciency for fluorophores that associate at high concentration. The concentration of the fluorophore was varied over four orders of magnitude, and at the highest concentrations, the fluorescence values had to be corrected for absorption by the fluorophore. 5 Van Geel et al. have developed a general expression for the observed intensity in molecular fluorescence spectrometry (2). They have divided the areas of the sample into the prefilter region (that part of the excitation beam that does not overlap with the emission beam), the post filter region (that part of the emission beam that does not overlap with the excitation beam), and the primary absorption/reabsorption region (where the two beams overlap). The expression assumes that the light is collimated and under those conditions is identical to the equation used by Ratzlaff (17) for a prefilter and post-filter length of zero. Instrumental Corrections Lutz has reported a simple cell shift instrument (21). This instrument shifts the cell along the diagonal of the cuvette. For the first measurement, both the excitation and emission pathlengths are short; for the second measurement the pathlengths are long. This method has the advantage of a simple expression for correction: ch F1(F1/F2)a where I“l is the short pathlength measurement, F2 is the long pathlength measurement, and a is given by 5/01 - 17), where ’1 is the short pathlength and 12 is the long pathlength. The worst case deviation of the absorption-corrected values was 3% with a primary or secondary absorbance as high as 2.0. Street has investigated the use of a mirrored sample cell compartment to enhance the observed fluorescence signal, and to diagnose and correct inner filter effects (22-25). Measurements are made with and without the mirrors covered. The ratio of the signal obtained with the mirrors blocked to the difference between the signals is related to the measured primary and secondary absorbance by a quadratic polynomial. Under conditions of collimated excitation and point source/collimated emission, the equation 6 reduces to a straight line with a zero intercept. The author suggests the construction of a working curve for each chemical system, making this method much less general than the cell shift method. Applications Many of the methods of correcting for inner filter effects continue to see use in applications of fluorescence spectroscopy. The equation of Demas and Adamson (26), originally derived to correct for errors caused when quenchers are added to a solution containing a fluorophore, has continued to be the most popular mathematical method. It was used successfully by Haga et al., who studied the quenching of Ru(bipyrizine)32+ by amines in acetonitrile (27). Gamache et al. had less success with this method (28). They found that the equation could not adequately correct their data in a study of the quenching of Cu(2,9-diphenyl-1,10-phenanthrocene)2+ by various Cr(III) complexes. The workers had to resort to the more complex and less precise lifetime method to determine the quenching constants for these systems. This underscores the need for more theoretically sound methods of correction, such as the cell-shift method. This equation has also been used by Marcus et al. to correct data for inner filter effects (29). They studied the binding of benzo(a)pyrene to cytochrome P-450. Blewitt et al. used the equation of Kinka and Faulkner (30) to study the quenching of the fluorescence of the tryptophan residues in dyptheria toxin at low pH, which was ultimately correlated to the tertiary structure of the protein (31). Dithiothreitol was added to the solution to reduce disulfide linkages; corrections were required to account for its absorbance at the excitation wavelength. The cell shift instrument built by Novak (32) has been used to correct for inner filter errors in the study of the quenching of the chlorophyll fluorescence by herbicide ureas (33). The corrections suggested for use with the side/bottom fluorometer by Batke 7 (20) have been used by Vas and Batke in studies of the binding of 3-phosphoglycerate to 3-phosphoglycerate kinase using ANS as both the competitive binder and fluorescent indicator (34). Schroeder et 0!. continue to report applications of the instrument based on the design of Holland (35). The latest applications involve membrane studies using a number of membrane-bound fluorophores. Dehydroergosterol has been used to study the fluidity of cancer cell membranes (36) and the effect of the fluidity on the action of squalene carrier protein (37). Cholestatrienol was used in a similar study (38). The fluorophore 1,6- diphenylhexa-1,3,5-triene (DPH) has been used to study the effects of cationic and anionic anesthetics on the fluidity of the leaflets of plasma cell membrane and on the action of several membrane-bound enzymes (39), and on the uptake of 7-aminobutyric acid (GABA) by the GABA receptor (40). Miscellaneous Phenomena Related to the Inner Filter Efi‘ect O’Neil and Schulman have discovered a simple means of determining the quantum efficiency of a fluorophore by the relative method, using highly concentrated solutions and right angle geometry (41). They found that, for a pure fluorophore, the product of the absorptivity of a compound and the concentration at which its maximum fluorescence occurs (where the calibration curve bends over) is dependent only on the geometric properties of the fluorometer used. The quantum efi'iciency of the unknown can be calculated by: ‘0 g 08 (Fu/Fs) where O“ is the quantum efficiency of the unknown, (as is the quantum efficiency of the standard, Fu is the maximum fluorescence intensity of the unknown, and F s is the maximum fluorescence intensity of the standard. 8 Chu et al. have put the inner filter effect to good use in the design of a universal fluorescence detector for HPLC (42,43). The mobile phase contains a small amount of a fluorescent compound. As a compound is eluted, the steady-state fluorescence signal either decreases (if the compound does not fluoresce) or increases (if the compound fluoresces). In the case of a non-fluorescent eluent, one of the mechanisms responsible for the decrease in the fluorescence is the inner filter effect (2). ABSORPTION-CORRECTION FOR FLOWING SAMPLE STREAMS Most work in absorption-corrected fluorescence to date has involved instruments that use conventional one centimeter square cuvettes. While this the is most versatile method of making fluorescence measurements, it does have some drawbacks. For example, none of the instrumental corrections developed are compatible with flowing sample streams. This is unfortunate, because many techniques in which fluorescence measurements are made on a flowing stream - including continuous flow analysis, flow injection analysis, high performance liquid chromatography and stopped-flow kinetics - could benefit from absorption correction. In order to fill this need, an instrument was designed by Ratzlafi to make primary absorption-corrected fluorescence measurements in real time (17,18). The sample cell is a cylindrical flow celL and light is conducted to and from the cell by a bifurcated fiber optic bundle. The instrument has three detectors, and simultaneously monitors source intensity, front surface fluorescence intensity and transmittance. A mathematical model was developed that corrects the observed intensity based on the value of the primary absorbance. The model developed by Ratzlaff, which is detailed in Chapter 2 of this dissertation, contains a term in which the primary and secondary absorbance are added together. By switching the excitation and transmittance monochromators to the emission 9 wavelength, and measuring the absorbance at this wavelength, it was possible to correct the fluorescence emission for secondary absorbance errors as well. However, the instrument was now no longer able to obtain the corrections in real time. Furthermore, to operate in near real time, it would have been necessary to slew the monochromators rapidly between the excitation and emission wavelengths, and either change the setting of the fluorescence monochromator or gate the aperture to the emission photomultiplier tube to protect it from the scattered excitation beam. The wavelength setting imprecision was thought to be a major source of error in the corrected values. This inability to correct for secondary absorption errors in real time was felt to be a major shortcoming of Ratzlaff’s instrument. The next three chapters describe an instrument and methodology that allow one to make fluorescence measurements corrected for both primary and secondary absorption errors in real time on a flowing sample stream. First, a fluorescence intensity expression that relates the measured fluorescence to the sample’s absorbance at the excitation and emission wavelengths is derived. Then the instrumentation and methodology are described, and experimental results are presented. Finally, the limitations of the current instrument are described. Some suggestions for its improvement are given, as well as suggestions for other ways of making absorption-corrected fluorescence measurements on a flowing sample stream. 990.495» 10. 11. l2. 13. 14. 15. 16. 17. 18. 19. 10 CHAPTER 1 REFERENCES Schulman, S.G., ed., Molecular Luminescence Spectroscopy, John Wiley and Sons: New York, 1985. Van Geel, F.; Voightman, E.; Winefordner, J.D. Appl. Spectrosc. 1984, 38, 228. Mielenz, K.D., ed., Optical Radiation Measurements, Volume 3: Measurement of Photoluminescence, Academic Press: New York, 1982. Lakowitz, J.R. Principles of Fluorescence Spectroscopy, Plenum Press: New York, 1983. Harkonen, M.; Adlercreutz, H. Trends Anal Chem. 1983, 2, 176. Holland, J.F.; Teets, R.E.; Timnick, A. Anal. Chem. 1973, 45, 145. Holland, J.F.; Teets, R.E.; Kelly, P.M.; Timnick, A. Anal. Chem 1977, 49, 706. Holland, J.F. Ph. D. Dissertation, Michigan State University, 1971. Christmann, D.R.; Crouch, S.R.; Holland, J.F.; Timnick, A. Anal Chem. 1980, 52, 291. Christmann, D.R.; Crouch, S.R.; Timnick, A. Anal. Chem 1981, 53, 276. Christmann, D.R.; Crouch, S.R.; Timnick, A. Anal. Chem. 1981, 53, 2040. Christmann, D.R Ph.D. Dissertation, Michigan State University, East Lansing, MI, 1980. Adamsons, K.; Timnick, A.; Holland, J.F.; Sell, J .E. Anal. Chem. 1982, 54, 2186. Adamsons, K.; Sell, J.E.; Holland, J.F.; Timnick, A. Am. Lab. 1984, 16, 16. Adamsons, K. M.S. Dissertation, Michigan State University, East Lansing, MI, 1982. Adamsons, K. Ph.D. Dissertation, Michigan State University, East Lansing, MI, 1985. Ratzlaff, E.H.; Harfmann, R.G.; Crouch, S.R. Anal. Chem. 1984, 56, 342. Ratzlaff, E.H. Ph.D. Dissertation, Michigan State University, East Lansing, MI, 1982. . Wiechelman, K. Am. Lab. 1986, 18, 49. 20. 21. i3 26. 27. 28. 29. 30. 31. 32. 33. 35. 36. 37. 38. 39. 40. 41. 42. 43. 1 1 Batke, J. Anal. Biochem 1982, 121, 123. Lutz, H-P.; Luisi, P.L. Helv. Chim Acta 1983, 66, 1929. Street, K.W.; Singh, A. Anal. Lett. 1985, 18, 529. Street, K.W. Analyst (London) 1985, 110, 1169. Street, K.W. Analyst (London) 1987, 112, 167. Street, KW. Analyst (London) 1987, 112, 921. Demas, J.N.; Adamson, AW. 1. Am Chem. Soc. 1973, 95, 5159. Haga, M.; Dodsworth, E.S.; Eryavek, (3.; Seymour,P.; Lever, A.B.P. Inorg. (Diem 1985,24, 1901. Gamache, R.E. Jr.; Rader, R.A.; McMillan, DR. 1. Am. Chem Soc. 1985, 107, 1141. Marcus, C.B.;T\1rner, C.R.; Jefcoate, C.R. Biochemistry 1984, 24, 5115. Kinka, G.W.; Faulkner, L.R. J. Am. Chem. Soc. 1976, 98, 3897. Blewitt, M.G.; Chung, L.A.; London, E. Biochemistry 1984, 24, 5458. Novak, A. Collect. Czech. Chem Commun. 1978, 43, 4869. Kaplanov, N. Photosynthesis 1985, I9, 221. Vas, M.; Batke, J. Eur. J. Biochem 1984, 139, 115. ' Schroeder, F.; Holland, J.F.; Vagelos, as. J. Biol Chem 1976, 251, 6739. Fischer, R.T.; Cowlen, M.S.; Dempsey, M.E.; Schroeder, F. Biochemistry 1985, 24, 3322. Schroeder, F.; Dempsey, M.E.; Fischer, R.T. J. Biol Chem 1985, 260, 2904. Kier, A.B.; Sweet, W.D.; Cowlen, M.S.; Schroeder, F. Biochim. Biophys. Acta 1986, 861 , 287. Sweet, W.D.; Schroeder, F. Biochem. J. 1986, 239, 301. Roberts, E.; Liron, 2.; Wong, E.; Schroeder, F. Era). Neurol. 1985, 88, 13. O’Neal, J .S.; Schulman, S.G. Anal. Lett. 1986, 19, 495. Su, S.Y.; Jurgensen, A.; Bolton, D.; Winefordner, J.D. Anal Lett. 1981, 14A, 1. Su, S.Y.; Lai, E.P.; Winefordner, J .D. Anal. Lett. 1982, 15A, 439. CHAPTER 2: A FLUORESCENCE INTENSITY EXPRESSION In previous work in absorption-corrected fluorescence, the excitation and emission beams were collimated to simplify the fluorescence intensity expression and make it integrable, and to satisfy one of the requirements of Beer’s law (1). However, light that exits a fiber or bundle of fibers is not collimated, so the expressions derived in previous studies are not applicable to this study. In this chapter an expression for the optical transfer function of a pair of fibers is derived, and is compared to a semiempirical transfer function that is integrable and forms the basis of the correction equation used in Chapter 3. Review Mitchell and coworkers were among the first to use optical fibers to measure molecular fluorescence in solution. They found that a front surface fluorescence measurement with a bifurcated bundle offered superior immunity to inner filter effects compared to measurements made with conventional 90° fluorometers (2). In designing a flowcell for their filter fluorometer, they studied the effects of the focal length of the excitation optics and the cell length on the observed fluorescence signal (3). Recently, the emphasis has shifted from bundles of fibers to single fibers for fluorescence measurements (4). Although much of the work involves the use of fluorescent reagents immobilized on fibers, there have been several reports of measurements of the fluorescence of discrete volumes of solution with single fibers (5-8), so the optical properties of fibers are still a concern. There have also been reports of the use of a pair of fibers to measure Raman scattering in solution (9,10). These researchers have investigated ways of maximizing the observed Raman scattering signal by using fibers with different diameters and numerical 12 13 apertures (defined below) (11,12) and by varying the angle formed by the fiber pair (12). In both studies mentioned above, the authors started with equations describing the optical properties of fibers and developed expressions for the observed signal that were numerically integrated. The model described below, although similar to those mentioned above, was derived independently of them and is analytically integrable. Fiber Optics An optical fiber consists of a core and a cladding with different indices of refraction. As shown in Figure 2-1, light that enters the fiber such that it strikes the fiber face at an angle of less than 0 undergoes total internal reflection and propagates along the fiber. If the angle of incidence is greater than a, the ray enters the cladding and leaves the fiber. Based on this simple model, it would be expected that the light leaving the end of an optical fiber is contained in a cone whose walls make an angle 0 with the vector normal to the fiber face. Furthermore, any rays not wholly contained within this same cone would not propagate along the fiber if incident on the fiber face. The sine of the angle defined by the wall of the cone and the vector normal to the fiber face is called the numerical aperture (n.a.) of the fiber, and is a function of the wavelength of light, and the refractive indices of the core, cladding, and external medium. Numerical apertures are usually reported for an external medium of air. Consider two fibers placed side by side, both of core radius r, and cladding thickness c, so that they have a total radius of R = r + c. At any distance 2 from the face of the fibers, the light emitted by one fiber forms a circle of radius B(2) 8 r + z - t, where t a tan (sin'1 na. ). The base of the acceptance cone of the other fiber is a circle with the same diameter, with its center located a distance of 2R away from the center of the other circle. Those two circles will not overlap until they are both of radius R, or in other words, until 14 >2 toEEmcob .oeeo 8:838» coca 38:5 .2 use..— \ OLOU poutEmcoblcoc 022620 15 r + c - r + z - t, or z = c / t. When they do begin to overlap, the overlap region is roughly elliptical, as shown in Figure 2-2. This region is symmetric about both the x and y axes; its area is given by 4 times the area of either circle evaluated from its extreme x value to where it intersects with the other circle. Adhering to the conventions of Figure 2-2, and using the equation for the circle on the left, the area of the overlap region is given by: B s) area(z)=4 [Bz(z)-xz]1/2dz R =2[(n/2-sin-1fimafin-Rmhzynzwz1 (2-1) Asz _. co, area(z) —- 1122, as expected. A Fluorescence Intensity Expression Assume that these fibers are immersed in a homogeneous medium with a linear absorption coefficient (absorbance/pathlength) at the excitation wavelength and emission wavelength of ‘ex and ‘em respectively. The medium contains a fluorescent compound with a linear absorption coefficient of Ecxf and a quantum efficiency of Q. In a plane located a distance 2 from the surface of the excitation fiber, the power at any point can be approximated by Pd = POIO"‘°"Z / 82(2), where P0 is the radiative power at the face of the excitation fiber. The fluorescence emitted by an infinitely small volume element at this point is PdQecxf dx dy dz. The emission is assumed to be isotropic. A certain solid angle of the emitted light is collected by the emission fiber; it is approximated by r2 / «(R2 + 22), where (R2 + 2:2)1/2 is the distance from the center of the overlap area to the center of the emission fiber. Before reaching the emission fiber, the emission is attenuated by absorption by a factor of 10““2. The power that enters the emission fiber is then given by: 16 cladding core overlap rodiu s=r+ zt ‘3") (b) x=0 formula of circle: F(B' (2)-X’)"’ Figure 2-2. Overlap region of light emitted from and collected by two adjacent optical fibers: (a) schematic diagram, (b) head-on View. 17 2 -8 z -8 z 1‘ P Tm} as: 10 ex 10 "n —2———-d:cdydz f 0 x B (z)1r(R2+zz) b .. f ’(eex ‘l’ 8cm) z r2 dz- (1 -130 Q Ens/10 [f—B‘I'Yz) 110224-22) dy z where b is the length of the cell. The double integral consists only of geometric quantities which relate to how much excitation radiation is reaching a point in the cell or how much light emitted from a point is collected by the emission fiber. Since none of the variables in the integral depend cm: or y, Equation 2-1 for the overlap area derived in the previous section can be combined with the equation above. The result, after converting to natural exponentials is: b 2 _ f -20303 (8e: + £2111): \ 1‘ Pf “Po Q 8e: 3]- e [area(z, 132(2) n(R2+zz) ] dz (2-2) Hereafter, this equation will be called the theoretical model. Summary of Previous Work If we compare Equation 2-2 to a similar equation derived by Ratzlaff (13), it is seen that the terms in brackets following the exponential term are identical to Ratzlaff’s transfer function, T(z), for which he proposed no theoretical expression. Instead, he assumed that there is an exponential rise of the transfer function near the face of the bundle, mostly due to the increase in overlap area of the cones of excitation and emission. A mathematical simulation showed that there is an erponential decrease in the transfer function at greater distances, due to the drop ofi in irradiated power and the decreased solid angle of collection. Based on those assumptions, the following transfer function was proposed: T(z)=G(e‘Hz-e'KZ) (2-3) 18 Hereafter, this will be called the experimental model. This transfer function has the important advantage that it is easily integrated after substitution into equation 2-2 for the term in brackets, to yield the following fluorescence intensity equation: _ f 1 _ -(H+2.303(sex+cem))b “)0 Q a" G[ H+2.303(£,,+ 8m) (1 e ) (2-4) 1 ( l_e-(x+2.303(e,,+ em))b)] K+2.303(£,,+£,m) The values of H and K were determined experimentally by measuring the front surface fluorescence and primary absorbance of a series of solutions of quinine sulfate (OS) (14). The resulting values were regressed on equation 2-4, with ‘ex equal to ‘exf and Gem: 0. The values reported by Ratzlaff are H = 1.108 :t 0.052, K = 9.60 :t 0.33. Comparison of Theoretical and Experimental Models The theoretical and experimental transfer functions are plotted in Figure 2-3, with the ordinate for the experimental transfer function on the bottom and the ordinate for the theoretical model on the top. Both curves are normalized to a maximum value of 1.0. Parameters for the theoretical model were chosen which approximate those of the equipment used in the experiments described in Chapter 3: r = core diameter = 116 um c = cladding thickness = 17 um na. 8 0.128 The value of the numerical aperture is not that of the fibers; it is the numerical aperture of the monochromators. The numerical aperture of an optical system is equal to the z o E z E.’ a: 11’ i 0.0 Figure 2-3. 19 * 1 ' 1 ' 1 ' 1 ' 0.2 0.4 0.0 0.0 1.0 DISTANCE FROM FIBER FACE (cm) Comparison of experimental and theoretical transfer functions. 20 smallest numerical aperture of all the components of the system. The monochromators are f/3; the numerical aperture is related to the f/no. by na. = (2-f/no.)'l, which gives a na. of 0.167 for the monochromators, compared to ca. 0.3 for the fibers. Since the ends of the fibers are immersed in aqueous solution with a refractive index of ca. 1.3, the effective na. of the system is 0.167 / 1.3 = 0.128. It can be seen that the theoretical model provides a close match to the shape of the experimentally determined transfer function. However, the difference in scale required to represent each on the same graph is disconcerting. The theoretical model indicates that the transfer function is zero until 2 > c / t, and also indicates that it peaks much closer to the fiber face. The first difference is easy to explain. It is assumed in the theoretical model that the optical power in the cladding is zero, but this is not true. If an electric field model, rather than the simple ray tracing model, is used to describe the propagation of light through an optical fiber, it is found that there is an ‘evanescent field’ in the cladding that is nonzero. This evanescent field produces an overlap area between the two fibers infinitely close to the face, as has been demonstrated by light scattering experiments (2). The difference between the locations of the peak maxima of the two transfer functions is more difficult to explain. The difference is probably a result of the omission of two factors in the theoretical model: (1) the fact that a fiber optic bundle was used in the experiments, whereas a fiber pair was employed in the model and (2) the effect of the cell walls was not considered. With numerical apertures near those used in the model, bundling would not shift the location of the peak maximum. This is because the maximum occurs at a distance less than that required to cause overlap between an excitation fiber and an emission fiber located at a distance of 2R. In a randomized 21 bundle, an emission fiber could be found at any distance from an excitation fiber, but this probabilistic effect is diffith to incorporate into the model. The presence of cell walls has two possible effects. If the walls are totally non- reflective, the overlap of the area between the two fibers is limited to 1R 3, where R c is the cell radius. This would cause a more rapid drop off of the transfer function at distances greater than where the top of the overlap area first contacts the cell wall; however, this would not effect the location of the transfer function maximum. If the walls did reflect some energy, which is likely, it would cause the incident power in the overlap area to be greater than estimated in Equation 2-2. Also, a larger solid angle of emission would be collected. This increase in power would be caused by the light which otherwise would have been lost reflecting off the cell wall. However, if large amounts of light were reflected off the walls, the absorption pathlength assumptions used to obtain the integrated fluorescence expression would not be correct. Two assumptions that were made in developing the theoretical model may also have contributed to the difference in the locations of the transfer function maxima. It was assumed that the power is constant everywhere in a plane at a given distance from the fiber face. A better approximation is a two dimensional Gaussian power distribution with a full width at half maximum defined by the numerical aperture (3). The other approximation that was made is that each point in the overlap plane is the same distance from the emission fiber as the center of the overlap area. Both assumptions were made so that the transfer function could be integrated analytically in the xy plane; without these assumptions numerical integration is required. Plaza et al. (12) compared the results of numerical integration both with and without the simplifying assumptions and found that the assumptions had little effect on the results. 22 Without abandoning the model derived here, there are two factors that could be included that would cause the maximum of the fiber transfer function to move farther out into the cell: an increase the estimate Of the fiber diameter and a decrease in the estimate of the numerical aperture. The effects of these two changes are shown in Figure 2-4. The first of these could be considered as a way of taking the bundling of the fibers into account. The second could be considered as a simple means to account for reflections from the cell walls. Conclusions The model described in this chapter has given the empirical transfer function of Ratzlaff (13) a more solid theoretical foundation. Because Ratzlaff’s transfer function has proven itself experimentally, and because the newly proposed equation is so unwieldy, the corrections in this dissertation are based on Equation 2-4, which was originally derived by Ratzlaff. The coefficients in this expression have already been proven experimentally to be independent of wavelength over a short range. This is expected, because the numerical aperture of the optical system is determined by the monochromators (wavelength-independent numerical apertures) rather than the fibers (wavelength-dependent numerical apertures). How much the coefiicients depend on the refractive index of the solution remains to be seen (15), but the dependence is expected tobe large. This is unfortunate, because it means that a new set of coefficients will have to be determined experimentally for each solvent in which fluorescence is measured, unless an equation can be found that relates the coefficients to the refractive index of the solution. Perhaps such an equation could be derived from the model presented here. TRANSFER FUNCTION TRANSFER FUNCTION Figure 2-4. 1.0 T 00" O 00‘ - 0.2 _ 1.0 0.4 0.2 ” , \\ fl 1 ' l ' j ' T ' I 0.0 0.2 0.4 0.0 0.0 1 .0 DISTANCE FROM FIBER FACE (cm) . ' (b) fi ' U I I | I I s —' 0.0 ' 0.2 0.4 0.0 0.0 1.0 ousrmc: FROM FIBER FACE (cm) Efi'ect of changes in the parameters used in the theoretical model: (a) change in numerical aperture (1 - 4 = numerical a rature of .1, .05, .025, .0125), (b) change in fiber radius (1 - 4 -- ber radius of .1, .3, .9, 2.7 mm). 9919933!" 9° 10. 11. 12. 13. 14. 15. 24 CHAPTER 2 REFERENCES Christmann, D.R.; Crouch, S.R.; Holland, J .F.; Timnick, A. Anal. Chem 1980, 52, 291. Mitchell, D.G.; Garden, J .S.; Aldous, KM. Anal. Chem 1976, 48, 2275. Smith, R.M.; Jackson, K.W.; Aldous, K.M. Anal Chem 1977, 49, 2051. Wolfbeis, O.S. Fresnius Z. Anal. Chem 1986, 325, 387. Vurek, G.G. Anal Chem 1982, 54, 840. Sepaniak, Ml; Tromberg, BJ., Eastham, J.F. Clin. Chem. 1983, 29, 1678. Renault, G.; Raynal, E.; Sinet, M.; Muffat-Joly, M.; Berthier, J .; Cornillault, J .; Godard, B.; Pocidalo, J. Am J. Physiol 1984, 246, H491. Chudyk, W.A.; Carrabba, M.M.; Kenny, J.E. Anal. Chem 1985, 57, 1237. McCreery, R.L.; Fleischmann, M.; Hendra, P. Anal. Chem 1983, 55, 148. Dao, N.O.; Plaza, P. Analusis 1986, 14, 119. Schwab, S.D.; McCreey, R.L. Anal Chem 1984, 56, 2199. Plaza, P.; Dao, N.O.; Jouan, M.; Fevrier, H.; Saisse, H. Appl Opt. 1986, 25, 3448. Ratzlaff, E.H. Ph.D. Dissertation, Michigan State University, East Lansing, MI, 1982, Equation 3-4, p. 44. Ratzlaff, E.H.; Harfmann, R.G.; Crouch, S.R. Anal Chem 1984, 56, 342. Victor, M.A.; Ph.D. research in progress, Michigan State University, 1987. CHAPTER 3: EXPERIMENTS IN ABSORPTION-CORRECTED FLUORESCENCE In this chapter the instrumentation and methodology for making fluorescence measurements that are corrected for both primary and secondary inner filter effects are described. The system is an extension of a previous system described by Ratzlaff (1,2) that corrects only primary inner filter errors in real time. THEORY An intuitive description of the phenomena underlying the correction method that was used in this work is explained below. Assume that monochromatic light enters a cylinder at the front of the cylinder as shown in Figure 3-1. The cylinder is assumed to contain a homogeneous distribution of a fluorescent material. If the light entering the cell is collimated, and is not attenuated significantly by absorption, then the right angle fluorescence measured at either end of the cell would be the same. However, if some of the light is absoer by the medium in the cell, the rear of the cell is less efficiently excited than the front. Since fluorescence is emitted isotropically, the front of the cell would present a larger solid angle for collection of emission than the rear, and the ratio of the fluorescence measured at the front to that measured at the rear would be less than one. Now imagine that the sample cylinder does not absorb the incident radiation, but does absorb the emitted radiation. In this case the fluorescence signals at both ends will decrease by the same magnitude. This is because the cylinder is uniformly illuminated, so the average plane of excitation is in the center of the cell, and the pathlength to either side is the same. Finally, consider the case where there is absorption at both the excitation and emission wavelengths. As before, the front is more efficiently illuminated than the rear, and on the basis of the solid angle of collection, the fluorescence signal at 25 26 n _ .388 =8 .8555 .3 use..— co.mm_Ee 385m .60.. Lo» 505.509 m100 times between 350 and 450 nm at pH 8 1, but their small absorptivities mean that very high concentrations (0.2 M) are required. At such high concentrations they were found to quench OS fluorescence. Another compound that was considered was N ,N-diethylparanitrosoaniline. Its synthesis was attempted, but the synthesis required the use of nitrite ion, which has a high absorptivity at 350 nm and is difficult to remove by recrystallization. Other compounds can be added to oxidized the excess nitrite to nitrate, but they absorb at 350 nm. The compound that was found to be most useful is sodium fluorescein. Although its absorptivity at 428 nm (A is only ca. 15 times that at 345 nm (1min), it is stable at low max) pH and its use as a secondary absorber is well documented (15). Although sodium fluorescein is fluorescent, its quantum yield is quite low at low pH. 33 ' EXPERIMENTAL DETAILS Reagents All stock solutions and dilutions were made 0.1 H in HClO4, which was prepared by diluting 9.0 mL of 72% HCIO4 (21) to 1.00 liter with distilled water. All solid reagents were used without purification. Solutions containing quinine sulfate (OS) (22) were made from a stock solution Of 1.00 mM OS. Solutions containing sodium fluorescein (SF) (23) were made from a stock solution Of 8.00 mM SF. A series of solutions of each combination of 0, 250, 500, 750, 1000, 1250 and 1500 11L of OS stock solution and 0, 125, 250, 375, 500, 625, 750, 875, and 1000 11L of SF stock solution was made and diluted to 10 mL in volumetric flasks with 0.1 N, HClO4. Procedure All measurements were made with solution flowing continuously through the sample cell at a rate Of 0.8 mL min'l. For each channel, 30 analog-tO-digital conversions were averaged and stored to create approximately a 1 second time constant. A blank signal was collected with 0.1 H HCIO4 flowing through the sample cell for 2 minutes. Then, the series of solutions containing 0 11L OS and 0 to 1000 11L SF was flawed through the cell, allowing 2 minutes for each solution, followed by 5 minutes of 0.1 H HClO4 blank. The source was then covered and the dark signal for each channel recorded. This procedure was repeated for each concentration of OS. The output of the four detectors for a typical experiment is shown in Figure 3-4. For each channel, the middle 80 seconds of the 120 second measurement for each solution were averaged together to produce one value. From this value, the dark signal for that channel was subtracted. Finally, the fluorescence channels were adjusted for source fluctuations and blank corrected by dividing by the reference channel 34 0.1 - (a) . 0.0 - L— 1.2 _, (b) 0.0 - 8 > 5.0 - (0) 0.0 - ——i 5.0 q (d) 0’0 T 1 ‘ I l 0 s 10 15 20 25 m: (mums) Figure 34. Output of the four detectors for 500 14M qUinine sulfate (OS), 0 to 910 um bromocresol purple (BP): (a) reference channel, (b) transmittance channel, (c) front surface fluorescence channel, d rear surface fluorescence channel. 35 measurement collected during the blank period, and subtracting the blank signal. For each concentration of OS and SF, the blank that was subtracted was the fluorescence measurement for the corresponding solution containing no OS but the same amount of SF. This was necessary because the blank signal was found to decrease with increasing SF concentration, indicating the large magnitude of stray light. The stray light problem will be discussed in Chapter 4. The transmittance signal was corrected for source fluctuations by dividing by the value of the reference channel measurement for that time period, then multiplying times the reference value for the blank signal. Transmittance was then calculated as (corrected transmittance) / (blank transmittance), and absorbance as the negative base 10 logarithm of the transmittance. The absorbance of each solution at 353 nm and at 438 nm was measured on a separate spectrophotometer (24). RESULTS AND DISCUSSION The ratio Of front surface fluorescence to rear surface fluorescence for each Of the 42 solutions containing OS is shown in Figure 3-5. As predicted, the ratio changes little with secondary absorbance when the primary absorbance is low (symbol= 1) and changes greatly when the primary absorbance is high (symbol=6). It can also be seen that the ratio changes little with primary absorbance when the secondary absorbance is low (left side of plot) but changes notably when the secondary absorbance is high (right side of plot). In order to use the fluorescence ratio as a way to estimate the secondary absorbance, a mathematical relationship between the ratio and the primary and secondary absorbance was sought. This was done by assuming the primary absorbance to be constant for each set of data in Figure 3-5, and fitting each data set to an equation that gives the ratio as a function of the secondary absorbance for that value of the primary 36 [SF]. W 0 10 20 30 40 50 60 70 so 13.0 1 1 1 1 1 1 1 s 7.0 - g 5.0 - < 1: 11.1 5.0 - o z I.1..I o K} 4.0 - a: o E 3.0 - 20 - ./ g M ‘-° n l l 1 l 1 1 1 1 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 ABSORBANCE AT 438 nM Figure 3-5. Fluorescence ratio versus absorbance at 438 nm. 1 = 250 um 1n1nesulfate (,QS 2= 500nm OS, 3= 750 #1511 OS, 4: 1000 um S, 5.. 1250uMQ,6=1500uMOS. 37 absorbance. A simplex fitting program written by Peter Wentzell based on the algorithm of Nedler and Mead (25) was used to determine the best fitting exponential equations, while a least squares polynomial fitting program (26) was used to determine the best fitting polynomial equations. In choosing an equation to fit to the data in Figure 3-5, two criteria were important: 1) the equation used to fit the data should be monotonic over the interval considered here, since the theory predicts that sort of behavior and 2) the parameters generated by the fit should themselves be monotonic with respect to the primary absorbance to reflect what is seen in Figure 3-4, and to produce a reasonable fit of the few resulting estimates of the parameters by an equation. Two models with three parameters were tried: ratio = a(A”)2+bA”+c, and ratio = aebA"+c, where A" is the secondary absorbance and a, b, and c are the variable parameters. Both functions were found to give parameters which were not monotonic. Since a two parameter polynomial fit is a straight line, and that would Obviously not work well, a fit of that equation was not attempted. Instead, two different two parameter exponential models were tried: ratio = aebA", and ratio 2 ea" + b. The former was found to be a better fit, and gave monotonic parameters with respect to the primary absorbance. These best fit equations are the lines drawn in Figure 3-5, and the agreement is good. The parameters a and b produced by the model are shown in Figures 3-6 and 3-7. These, too must be fit to an equation to allow interpolation between experimental points. Parameter a appears to rise exponentially, so both of the two parameter exponential models were tried; neither seemed to produce a reasonable fit. However the three parameter exponential model, a a few, + h did produce reasonable results, with f = 0.00386 , g = 2.02 and h = 1.20 . The second parameter, b in Figure 3-7, did not appear PARAMETER a 1.45 .1.40 1 .35 1 .30 1.25 1.20 38 0.0 0.4 0.0 1.2 1.0 . 2.0 ABSORBANCE AT 353 nm Figure 3-6. Parameter a versus absorbance at 353 nm. PARAMETER b 1 .25 - 0.75 -* 0.50 - 0.25 - 39 0.00 0.0 Figure 3.7. I I I l I I 0.4 0.8 1.2 ABSORBANCE AT 353 nm Parameter 6 versus absorbance at 353 nm. 1.6 2.0 40 to have any particular shape to it, so it was fit by a straight line b = jA' + k, where j = 0.551 and k = 0.144. It is possible to reverse the procedure and estimate the secondary absorbance of each solution. The algorithm is as follows: 1) With the measured values Of the primary absorbance and f, g, and h, calculate a. 2) With the measured value of the primary absorbance and j and k, calculate b. 3) With the measured value of the front/ rear surface fluorescence ratio R, and a and b, calculate the secondary absorbance asA" = log (R / a) / b. The estimates of the secondary absorbances of these solutions calculated by the procedure above are shown in Figures 3-8, 3-9 and 3-10 and are summarized in Tables 3-1 and 3-2. In these figures and tables, the estimated secondary absorbance is compared with the value measured on a separate spectrophotometer (20). The residuals in Table 3- 1 are the differences between the measured and estimated secondary absorbances. The variance of linear regression in Table 3-2 is the root-mean-square of the differences between the calculated or measured secondary absorbances and the line that best fits the points. , The results show that the estimated secondary absorbance is not as linear with concentration as the measured secondary absorbance, as would be expected. The average difference between the estimated and measured values is ca. 0.03 absorbance units. At all concentrations of OS, the relative residual is largest at low concentrations of SF, and the estimates are always lower than the measured absorbances. The consequences of this will be evident when the corrected fluorescence values are shown. 41 (a) ABSORBANCE AT 438 nm 1.6 -‘ (b) 1.4 - 1.2 '- 1.0 ~ 0.0 -‘ ABSORBANCE AT 438 nm 0.4 - 0.2 '- o.0 -1 [5"]. IN Figure 3-8. Absorbance at 438 nm (1 I: calculated, 2 8 measured) versus sodium fluorescein SF) concentration: (a) 250 uM Quinine sulfate (05). (b) 500 14M 0 . 42 A m V ABSORBANCE AT 438 nm (b) E C ,9, i’ ’2 la) O i 8 9 I I I I I I I I 0 10 20 30 40 50 00 70 00 [SF]. W Figure 3-9. Absorbance at 438 nm (1 2 calculated, 2 = measured) versus sodium fluorescein (SF) concentration: (a) 750 um quinine sulfate 43 A 33 ABSORBANCE AT 438 nm (b) Aesonamcs AT 438 nm [5F]. I‘M Figure 3-10. Absorbance at 438 nm (1 = calculated, 2 = measured) versus sodium fluorescein (SF) concentration: (a) 1250 pm Quinine sulfate (OS), (b) 1500 14M OS. TABLE 3-1. COMPARISON OF MEASURED AND CALCULATED SECONDARY ABSORBAN CE IOS|,10'6M AVE. RESIDUAL MAX. RESIDUAL MIN. RESIDUAL 250 0.018 0.04 0.005 500 0.024 0.07 0.002 750 0.045 0.07 0.003 1000 0.040 0.09 0.02 1250 0.027 0.04 0.008 1500 0.022 0.1 0.001 TABLE 3.2. COMPARISON OF LINEAR REGRESSION RESULTS FOR MEASURED AND CALCULATED SECONDARY ABSORBANCE VARIANCE OF R REGRESSION (x 10 ) MEASURED CALCULATED SECONDARY SECONDARY [OS]I 10'6M ABSORBANCE ABSORBANCE 250 7.0 16 500 6.5 19 750 9.5 23 1000 9.2 20 1250 3.7 14 1500 6.5 21 45 The values of the secOndary absorbance, along with the measured values of the primary absorbance, can now be used to correct the measured front surface fluorescence values for absorption effects. The equation was derived by Ratzlaff (24); its derivation, and a model which justifies the use of a formerly empirical transfer function, is presented in Chapter 2. The equation is: F: = F” - C where Fc is the corrected fluorescence, F” is the measured front surface fluorescence and C is a correction factor which is given by: C = f,.,,°/fm where fm" and fm are equations which give the theoretical observed fluorescence signal in the absence and presence of absorption effects, respectively. The value offIn is given by: f = l (1_e-(1.1os+2.303(c,,+5mm) m 1.1oa+2.303(a,,+£,m) l - -(0.s+2.303(e,,+€.m))b) 9.6+2.303(8,,+8,m) (l-e where ‘ex and ‘em are the linear absorption coefficients (absorbance/pathlength) for excitation and emission respectively, and b is the pathlength, which was experimentally determined to be 1.00 cm. The value offm‘ is given by the equation above with ‘ex and rem equal to zero. The results of the correction procedure are shown in Figures 3-1 1, 3- 12 and 3-13 and are summarized in Table 3-3. The range in Table 3-3 is the difference between the smallest and largest values in the data set. The corrected fluorescence values follow the same trend as the estimated secondary absorbance. For each concentration of OS, the second corrected fluorescence value is always lower than the first, which has no SF. This is due to the systematically low 46 TOTAL ABSORBANCE (353 nm + 438 nm) 0.50 0.75 1 .00 1 .25 1 .50 1.75 2.00 .0 l J J l J l l .W (a) ”j ' a... 3 3.4 8 8 8 8 3 8 [Sf]. 1N TOTAL ABSORBANCE (353 nm 4 438 nm) 0.75 1.00 1.25 1.50 1.75 200 2.25 J J l I l I 1 Ni: -¥ P ‘5 4N (b). - surmooscuaaanmx 8 1 I . o 3 8 8 8 3.1 g- 3. [SF]. m Figure 3-11. Front surface fluorescence intens312~)( 1= measured, 2 = corrected using calculated absorbance at 4 11m) versus sodium fluorescein concentration: (a) 250 pM quinine sulfate (OS), (b) 500 um OS. 47 TOTAL Aesoaamcs (355 nm + 430 nm) 1.00 1.25 1.50 1.75 2.00 2.25 2.50 l l l l J l l (a) 150: m PMTANOMCURRENT.M 8 l 30 Cl ° 1 I 1 I U I l 1 I 0 10 20 30 40 50 50 70 50 [57]. W TOTAL ABSORBANCE (353 nm 4' 458 nm) 1.25 1.50 1.75 2.00 2.25 2.50 2.75 5.00 240 1 1 1 1 J 1 1 1 (b) ‘ \ = -.- - - zoo -1 . . N E d g 150 " 120 - 5 '° ' ‘0 - o I I r I 1 1 I I r 0 10 20 50 40 50 .0 70 .0 [SF]. u“ Figure 3-12. Front surface fluorescence intensitgy (1 = measured, 2 = corrected using calculated absorbance at 43 nm) versus sodium fluorescein concentration: (a) 750 MA quinine sulfate (OS), (b) 1000 AM OS. (a) . 11A PMTANODE (b) Figure 3.13. , 11A 'PMTANODE 48 101111 1135012310105 (553 nm 4» 453 nm) 1.50 1.75 2.00 2.25 2.50 2.75 5.00 5.25 m J I I I J J I I 240 '- 200 '1 100 "I .1 120 -' m .1 2 .\.\‘\.\¥ ‘0 q ' A 1 ° 1 1 1 I 1 r 1 r r 10 20 50 4O 50 50 70 00 [SF]. W TOTAL ABSORBANCE (555 nm + 458 nm) 1.50 1.75 2.00 2.25 2.50 2.75 5.00 5.25 5.50 1 1 1 1 L 1 1 1 1 500 _ N 250 '- «1 200 -1 150 .1 100 - w - \ ° 1 1 1 1 1 1 1 1 1 0 10 20 50 40 50 00 70 00 isri. uM Front surface fluorescence intensi (1 = measured, 2 = corrected using calculated absorbance at 43 nm) versus sodium fluorescein concentration: (a) 1250 um quinine sulfate (OS), (b) 1500 11M 08. 49 TABLE 3-3. FLUORESCENCE VALUES 6 REL. [OS], 10' M “ MEAN ST. DEV. ST. DEV. RANGE 250, U 21.0 9.6 0.46 C 56.2 0.91 0.016 500, U 32.2. 13.4 0.42 C 105 1.8 0.017 750, U 39.1 14.6 0.37 C 158 2.5 0.016 1000, U 45.3 16.6 0.37 C 209 7.6 0.036 1250, U 47.4 15.2 0.32 C 263 2.6 0.010 1500, U 49.1 14.8 0.30 C 312 6.5 0.021 “ U = uncorrected, C = corrected ST. DEV. = standard deviation REL. ST. DEV. = relative standard deviation 29 2.6 41 5.6 44 8.0 51 27 46 8.8 45 24 COMPARISON OF UNCORRECTED AND CORRECTED REL. RANGE 1.4 0.046 1.3 0.053 1.1 0.050 1.1 0.13 0.97 0.033 0.91 0.078 50 estimate of the secondary absorbance for low concentrations of SF that was described earlier. Furthermore, the corrected fluorescence values peak near the middle of the concentration range, and drop Off again at high concentrations of SF. It may be that the correction equation is no longer usable at high absorbances. However, these data Show that the correction procedure is useful up to an absorbance of at least 3.25. Note that the relative standard deviation and relative range of the corrected values, as shown in Table 3-3, are 10 to 30 times lower than the uncorrected values. The corrected values have an average relative standard deviation of 1.9% and an average range of 6.5%. A Test of the Correction Procedure INTRODUCTION It is not surprising that parameters generated from fitting a curve to a set of experimental points could then be used to generate a set of curves which fit that same set well. In order for a reasonable test of the correction procedure to be made, it must work well when used on another set Of data generated under a different set of conditions. In order to accomplish this, a compound was sought which is stable at pH= 1, is not fluorescent and does not quench the fluorescence of the OS. Furthermore, it is important that the compound chosen have absorptivities at 345 and 438 run that are nearly equal, so that both primary and secondary absorbances could be varied at roughly the same rate. Alizarin red was tried first, but it quenched the fluorescence of OS. This was ultimately traced to the presence Of NaCl in the unpurified indicator. However, even after recrystallization from ethanol the alizarin red still quenched the OS fluorescence. The next choice, bromocresol purple (BP), was found to be nearly ideal. Its only drawback was its surfactant character, which made pipetting difficult. Figure 3-14 is a spectrum of 909 11M BP in 0.113 HClO4. 51 secs 2 3 5 CE 2&3 68885 2.. 83 Co 5.58% 5:98? EA Esme 2:5 Ikozmmm><3 com com 4 » 0.0 nan mme i/ V 8 S O H 8 V N 3 3 ON 52 WERIMENT’AL DETAILS Reagents Stock solutions of I-IClO4 and OS were used as described in the previous section. Solutions containing bromocresol purple (BP) (20) were made from a stock solution of 9.09mM BP. A series of solutions consisting of each combination of 0, 250, 500, 750 and 1000 14L OS stock solution and 0, 100, 200,..., 900 11L BP stock solution was made and diluted to 10 mL in a volumetric flask with 0.1 H HClO4. Procedure The procedure for measurement of these solutions was the same as that in the previous section. In addition, the absorbance of these solutions was measured at 438 nm by changing the setting of the excitation monochromator until the largest scattering signal was measured at the fluorescence channel, then the transmittance monochromator setting was varied until the largest transmittance Signal was measured. After these adjustments were made the absorbance was determined. RESULTS AND DISCUSSION Figures 3-15 and 3-16 Show the estimated secondary absorbance for each of the 40 solutions which contained OS. The results are summarized in Tables 3-4 and 3-5. These secondary absorbances were estimated with the parameters found for SF in the previous section and the measured values of the primary absorbance Obtained by this instrument. Except for the series of solutions containing 250 11M OS, the values of the average residuals are about the same as those found in the previous section. It is apparent that the ability of this method to predict the secondary absorbance of the solutions is not reliable at small primary absorbances, as was predicted. In fact, an absorbance of less than zero is estimated in several cases. The results shown in Figure 3-15a are particularly A m v ABSORBANCE AT 458 nm (b) ABSORBANCE AT 458 nm 1.60 -' 1.40 -' 1.20 -‘ 1.00 -' (LOO - (L00 '- €L40 '- IL20 - (L00 '- 53 -{L20 1.60 1 1.40 - 1.20 - 1.00 - (L00 - (L00 - (L40 '- IL20 - (L00 -‘ I I I I I I I _ I I I 100 200 500 400 500 600 700 .000 000 [3"]- IN -4L20 Figure 3.15. 1' I I I I I I . I I 100 200 500 400 500 000 700 000 000 [9"]- W fibsorbancel at 4318e 11mP 1 = calculated, 2 (=) néggsured) versus romocreso Sconcentration: a 11M uinine sulfate (OS), $311000 amp 0 1.50 -' 1.40 '1 A 33 1.20 -' 1.00 -' 0.00 -' 0.00 -' 0.40-I ‘ ABSORBANCE AT 458 nm 0.20 -‘ 0.00 -1 "”0 I 1 1 I I I 1 1 T I 0 100 200 300 400 500 500 700 500 000 [SP]. W 1.50 - (b) 1.40 - 1.20 - 1.00 4 0.50 «- 0.50 '- 0.40 -' ABSORBANCE AT 458 nm 0.20 - 0.00 -' Figure 3-16. Absorbance at 4318c nrr1P 1 = calculated, 2 = measured) versus bromocresol concentration: (a) 750 uM quinine sulfate (OS), )000(B 11M OS 55 TABLE 34. COMPARISON OF MEASURED AND CALCULATED SECONDARY ABSORBAN CE IQSI: 10'6M AVE. RESIDUAL MAX. RESIDUAL MIN. RESIDUAL 250 0.072 02 0.004 500 0.046 0.1 0.004 750 0.026 0.06 0.001 1000 0.013 0.04 0.002 TABLE 35. COMPARISON OF LINEAR REGRESSION RESULTS FOR MEASURED AND CALCULATED SECONDARY ABSORBANCE VARIANCE OF R REGRESSION (x 10 ) MEASURED CALCULATED 6 SECONDARY SECONDARY [081,10 M ABSORBANCE ABSORBANCE 250 2.8 58 500 1.7 37 750 33 26 1000 1.4 44 56 discouraging. However, at high primary absorbances the agreement is respectable, as it is in at least one region of each Of the four data sets. With the measured value of the primary absorbance, and the estimated value of the secondary absorbance, the fluorescence values were corrected for inner filter effects as described in the previous section. The results of the correction are shown in Figures 3- 17 and 3-18, where these values are compared to the corrected values Obtained with the measured primary and secondary absorbances. The results are also summarized in Table 3-6. The corrected values obtained with the measured secondary absorbances, those labelled M in Table 3-6, are as good as the results Obtained with SF in the previous section. The average relative standard deviation for the four data sets is 1.9%, and the average relative range is 5.3%. Unfortunately, the corrected fluorescence values obtained with the estimated secondary absorbances, those labelled C in Table 3-6, are not as good, particularly with the series of solutions that contained 250 pM OS. For the corrected fluorescence values Obtained with the estimated secondary absorbances, the average relative Standard deviation is 3.3% and the average relative range is 10%, which is about half as good as the values obtained with the measured secondary absorbances. For these solutions, the corrections appear to be useful to a total absorbance of 2.5. Real-time Performance Figures 3- 19 and 3-20 show the real-time performance of the instrument in determining the absorption-corrected front surface fluorescence of a series of solutions containing 500 14M OS and 0 to 909 um BP. Figure 19a Shows the measured primary absorbance. The spikes in this trace mark the occurrence of a bubble flowing through the cell. The bubble is introduced when the sample tube is moved from one sample container to another, and serves as a marker between solutions. The difference in refractive indices (a) . M (b) www.m' Figure 3-17. Furnaces. 57 - 10m. Aesosamcz (353 nm + 458 nm) 0.25 0.50 0.75 1 .00 1 .25 1 .50 I .75 2.00 2.25 2.50 a I L I I I 4 I J I ‘ r, 7“: 4° -1 q a .1 a .1 10 ‘- ° I I I I 1 1 I I I 100 200 500 400 500 000 700 000 000 [8?]. W TOTAL ABSORBANCE (555 11111 + 458 nm) 0.50 0.75 1 .00 1 .25 1 .50 1.75 2.00 2.25 2.50 2.75 00 I I I I J I I I I I w u 04 -1 a - a - 10 '- ° I I I I I m I I —T 0 100 200 500 400 500 000 7” 000 000 [SP]. uM Fluorescence intensity (1 s uncorrected, 2 -= corrected usin calculated absorbance at 438 nm, 3 = corrected using measure) absorbance at 438 nm) versus bromocresol urp le concentration: (a) 250 11M quinine sulfate (OS), (b) 1100M O. 58 TOTAL ABSORBANCE (555 nm + 458 nm) 0.75 1 .00 1.25 1.50 1 .75 2.00 2.25 2.50 2.75 5.00 (a) I .0 I I I I I I I I I J W PMTANWECURRENT.IIA 8 In 20 - J ° I i I I l I I I I 0 100 200 500 400 500 000 700 000 000 [3"]- IN TOTAL ABSORBANCE (555 nm 4' 458 nm) (b) 1 .00 I .25 1 .50 1 .75 2.00 2.25 2.50 2.75 5.00 5.25 1 1 1 I I I I 1 1 1 I” -' M 150 "' E . E 120 - g .0 .- 5 '° - a - ° I I I I I I I I I 0 I” 2W 5W 4W 0” 000 100 000 ”0 [BP], nu Figure 3-18. Fluorescence intensity (1 - uncorrected, 2 1- corrected usin calculated absorbance at 438 nm, 3 = corrected using measure absorbance at 438 nm) versus bromocresol p 1e éBP) concentration: (a) 750 11M quinine sulfate (OS), (b) 1 14M S. 59 TABLE 3-6. FLUORESCENCE VALUES 6 REL. [OS], 10' M * MEAN ST. DEV. ST. DEV. RANGE 250, U 16.2 8.1 0.50 C 44.9 2.1 0.046 M 46.6 0.91 0.019 500, U 25.4 12 0.47 C 87.6 2.1 0.024 M 88.1 0.88 0.0099 750, U 31.0 13 0.42 C 130 3.0 0.023 M 130 1.6 0.012 1000, U 36.9 15 0.41 C 170 6.5 0.038 M 170 6.1 0.036 f U = uncorrected, C = corrected using the calculated 2 absorbance, and M = corrected using the measured 2 absorbance. ST. DEV. = standard deviation REL ST. DEV. = relative standard deviation 25 7.0 2.5 37 5.8 2.7 40 8.0 4.0 46 18 17 COMPARISON OF UNCORRECTED AND CORRECTED REL. RANGE 1.6 0.16 0.053 1.5 0.067 0.031 1.3 0.061 0.031 1.2 0.11 0.098 8 0.0 .— i 3 0.5 - g 6.. - a. 0.2 -1 (a) 0.0 -——i ' .L__ g .. a 1.2 - g 0.9 - é 1 . (b) ‘I’IME (MINUTES) 20 25 Figure 3-19. Real-time results: (a) measured primary absorbance, (b) estimated secondary absorbance. 61 so- ,. (a) 8 1°- 8 g..- 5 5... ...—J “ (b) 8 .. 8 :40- 5 05-20- J o l I I. I LI— _ 0 5 10 15 20 25 TIME (MINUTES) Figure 3-20. Real-time results: (a) front surface fluorescence, (b) corrected front surface fluorescence. - 62 between the aqueous sample and the air in the bubble causes most of the excitation beam to be reflected at the air/ solution interface, and thus only a small fraction of the beam reaches the other end of the cell. This has the same effect as a high sample absorbance. Figure 3-19b shows the estimated secondary absorbance, based on the measured primary absorbance and the ratio of the measured front surface fluorescence to the measured rear surface fluorescence (not shown). The spikes in this trace occur for the reason given above, but they are usually in the opposite direction as the spikes in the primary absorbance trace. For a given value of the fluorescence ratio, the higher the primary absorbance, the lower the estimated secondary absorbance, which explains why they should go in opposite directions. The large spike near 15 minutes is due to incorrect placement of the sample tube. This solution was the run twice as long as the others. Note the concentration linearity Of the estimated secondary absorbance. Figure 3-20a shows the measured front surface fluorescence, and Figure 3-20b shows the corrected front surface fluorescence. These figures reflect what was shown in Table 3-6: the corrected values have a much lower standard deviation, and the range of highest value to lowest value is smaller. The microcomputer that controls the instrument (16) is capable of displaying a corrected fluorescence value at the rate of 5 Hz, which is fast enough to accurately characterize continuous flow analysis (CFA) and flow injection analysis (FIA) peaks. The measured front surface fluorescence, shown in Figure 3-20a, is the least noisy of the four traces. The other three traces rely on the transmittance measurement, either directly, as in the case of the primary absorbance (Figure 3- 19a), or indirectly via a calculation, as in the case of the estimated secondary absorbance (Figure 3-19b) and the corrected front surface fluorescence (Figure 3-20b). The end result is that the corrected front surface fluorescence (Figure 3-20b) is noisier than the measured front surface 63 fluorescence (Figure 3-20a), primarily due to the uncertainty in the transmittance measurement. This weak link in the system will be addressed in the next chapter. 99°89???“ 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. CHAPTER 3 REFERENCES Ratzlaff, E,H.; Harfmann, R.G.; Crouch, S.R. Anal Chem. 1984, 56, 342. Ratzlafi', E.H. Ph.D. Dissertation, Michigan State University, East Lansing, MI, 1982. The Second Source, Duarte, CA. Model M303, PRA International, Inc., Oak Ridge, TN. Model H10, Instruments SA, Metuchen, NJ. Maxlight Optical Waveguides, Inc., Phoenix, AZ. Model MC1-02, PTR Optics Corp., Waltham, MA. Model A83210-M2, North American Philips Control Corp, Cheshire, CT. Model R1527, Hammamatsu Corp., Middlesex, NJ. Highlight Fiber Optics, Caldwell, ID. Model S780-5BO, Hammamatsu Corp., Middlesex, NJ. Patton, C.J. Ph.D. Dissertation, Michigan State University, East Lansing, MI, 1982. Model EU42A, Heath Corp., Benton Harbor, MI. Model 427, Keithley Instruments Inc., Cleveland, OH. Christmann, D.R., Ph.D. Dissertation, Michigan State University, East Lansing, MI, 1980. System 9000, IBM Instruments Inc., Danbury, CT. LSI-ll, Digital Equipment Corp., Maynard, MA. Cadillac Plastic and Chemical Co., Detroit, MI. Ismatec Model IP- 12, Cole-Parmer Instrument Company, Chicago, IL. Bishop, E., ed. Indicators, Pergamon Press: New York, 1972. Malincrodt, Inc., Paris, KY. N 9‘ 65 Aldrich Chemical Co., Inc., Milwalkee, WI. Eastman Kodak Company, Rochester, NY. Model Lambda 5, Perkin Elmer Corp., Norwalk, CT. Nedler, J.A.; Mead, R. ComputerJ. 1965, 7, 308. Johnson, K.J. Numerical Methods in Chemistry, Marcel Dekker, Inc.: New York, 1980. CHAPTER 4: PERSPECTIVES, CONCLUSIONS, AND SUGGESTIONS FOR FURTHER STUDY The instrument described in the previous chapter allows one to correct fluorescence measurements for inner filter efi'ects in real time. The performance of the instrument lies somewhere between standard fluorescence Spectrometers that use mathematical corrections of fluorescence measurements (low end) and the current state of the art in cell shift technology (1). In this chapter some basic limitations of the instrument are discussed, then some comparisons of the new instrument to the cell shift instrument are made. Finally, some suggestions for future work are proposed. LIMITATIONS OF THE CURRENT INSTRUMENT AND SUGGESTIONS FOR IMPROVEMENT One of the major limitations of the instrument thus far appears to be the ability to measure the primary absorbance. A photodiode was chosen as the transducer of the transmitted light to simplify the system by not requiring a high voltage power supply, but a photomultiplier tube (PMT) would have been a better choice. Not only are photomultipliers much more sensitive to ultraviolet light (UV) than UV enhanced photodiodes, but most PMT’s have peak sensitivity in the blue region of the spectrum. Photodiodes Show maximum sensitivity in the red, and thus are quite susceptible to stray light in the infrared. Due to the low UV sensitivity of the photodiode, a conversion factor of 1011 V/A was required to measure its current, which made this measurement extremely susceptible to noise. Since the front surface/ rear surface fluorometer incorporates a flow cell, the flow characteristics of the system are of importance. Because Ratzlaff reported that his instrument had poor wash characteristics (2), the wash time of the instrument described in this thesis was tested. A solution Of 100 11M OS was passed through the cell, followed 67 by a four-second bubble, followed by 0.1 N HClO4. The time profile of the rear surface fluorescence measurement after the bubble passed through the cell is shown in Figure 4- 1. Figure 4-1a shows that the half-life of cleaning the cell is roughly 1.25 minutes. Figure 4-1b shows that the measured fluorescence settles to within 20% of the steady-state blank value of 0.35 nA after about 20 minutes, which would make it difficult to exploit this instrument for the high throughput methods Of flow injection analysis (FIA) and air segmented continuous flow analysis (CFA). The wash characteristics of the instrument could be improved with a cell that is made out of quartz instead of Delrin, and is surrounded with some black material to satisfy the non-reflective cell wall requirements of the model described in Chapter 2. Quartz has the added advantage of being more inert chemically than Delrin. Unfortunately, use of a quartz flowcell is expected to cause two problems. The first is that some reflection is bound to occur between the solution/glass interface and the glass wall/black surface interface. The second is that each end would now consist of a fiber/air/glass/solution interface rather that a fiber/ solution interface, which would increase the amount of reflected light, as well as increase the stray light reaching the PMT’S. The reflections could be minimized by flooding the region between the fiber bundle and the glass wall of the cell with a liquid with a refractive index nearly equal to that of quartz. Although absorption-correction techniques can extend the upper end of the dynamic range of fluorescence measurements, the lower end is still important; after all, sensitivity is one of the reasons for choosing fluorescence spectrometry instead of another method. With a S / N criterion of 2, the detection limit of the front surface/ rear surface fluorometer was found to be 0.01 11M quinine sulfate (OS) , which is comparable to what is obtained by the cell shift instrument (3). The greatest source of noise was, as in Ratzlafi’s instrument, the source flicker noise superimposed on the high blank signal. PMT ANODE CURRENT. nA PMTANODECURRENT,IIA Figure 4-1. I I I 1 2 5 4 TIME (MINUTES) (9) TIME (umurss) Time profile of rear surface fluorescence measured while sample cell is rinsed with 0.1 N HClO4: (a) 0 to 4 minutes, (b) 0 to 25 minutes. 69 Figure 4-2 shows the front surface fluorescence Spectrum of the 0.1 N HClO4 blank, both with (Figure 4-2a) and without (Figure 4-2b) the bandpass filter in front of the excitation monochromator and the longpass filter in front of the PMT. Without the filters, most of the blank signal is due to the tail of the excitation peak at 353 nm. There does appear to be some additional contribution to the tail near 410 nm, 470 nm, and 540 nm. The peak near 540 nm is probably a reentry peak due to specular reflection of the scattered 353 nm excitation beam inside the emission monochromator. The peak near 410 nm is probably a Raman band of water. The peak near 470 nm can be seen more clearly in the spectrum obtained with the filters in place. It is the spectrum of the Xenon arc lamp that was used for excitation. There does not appear to be any fluorescence contributing to the blank signal. Further evidence that the large fluorescence blank signal was not due to adsorbed OS, as was previously suggested by Ratzlaff, was Obtained by measuring the fluorescence blank Signal from 1 M NaCl in water and 1 M NaOH in water. Neither blank signal was lower than that due to the 0.1 H HClO4. NaCl is a strong quencher of OS fluorescence (4), and the quantum efficiency of OS is low at high pH, so both solutions would be expected to suppress any contribution from adsorbed OS to the blank Signal. It is still possible that there was some contribution to the fluorescence blank from the fluorescence of impurities in the quartz fibers, in the epoxy that holds the fibers together, or in the Delrin flow cell, but this is considered unlikely. It would be possible to distinguish fluorescence from stray light if the lifetime of the blank could be determined. It is not difficult to understand why the blank signal at the rear of the cell should be so large: the source is pointed directly at it. It is more difficult to understand why the signal is so large at the front end of the cell. Part of it is undoubtedly due to reflections from the rear of the cell, but only about 0.2% of the incident light is expected to be reflected at a quartz/water interface. Thus, the stray light at the front is expected to be at 70 0.5 - (a) 3. 8 8 5 5 O. 0.0 v I I I ' l ' I 400 450 500 550 WAVELENGTH (nm) 0.02 '- (b) E E 8 3 O 5 ’é 0.00 -- - . 1 . I . r 400 450 500 550 WAVELENGTH (nm) Figure 4-2. Frorlit surface fluorescence blank: (2) filters removed, (b) filters in p ace. 71 most 0.002 times that at the rear. This assumption neglects the light reflected by the epoxy between the fibers. It is also possible that a portion of the stainless steel ferrule that surrounds the rear fiber bundle is aposed to the excitation beam, which would dramatically increase the amount Of light that is reflected back to the front from the rear of the cell. Broken fibers in the common end Of the bundle also contribute to stray light in the front surface measurement. Light that leaves a broken excitation fiber in the common leg of the bundle can enter a broken emission fiber. This problem of high blank signal is both a technological and a theoretical concern. Although it would be possible to remove more stray light with better monochromators and better filters, and thus reduce the blank Signal, the blank could never be made as low as that Obtained from a 90° fluorescence measurement in a cuvette. The predominant source of noise is the source flicker noise, which manifests itself both within the blank and the fluorescence Signal Source flicker noise and drift is normally minimized by dividing the fluorescence signal by the reference Signal, as was done in these experiments. Examination of the reference, transmittance and fluorescence signals revealed that the reference Signal could not be correlated with the others over short time periods, but only over periods of minutes or more. In fact, the reference channel Signal would Often fall when the fluorescence rose and the opposite also held true. This could have been due in part to the different time constants of the amplifiers used, but the major cause is more likely the spatial characteristics of the source flicker noise from the arc lamp. The reference signal is produced by a photodiode connected to a single fiber placed in the vicinity of the exit slit Of the excitation monochromator. Thus only a small part of the lamp image that formed the excitation beam is sampled. In a typical fluorescence spectrometer a fraction of the entire excitation beam is sampled, so spatial noise is averaged out by the reference detector (5). 72 Two modifications to improve the quality of reference correction are suggested. The first is to replace the single reference fiber with a bundle of fibers that are randomly positioned along the exit slit. The second solution is to split off a fraction of the source before the entrance slit of the excitation monochromator. This is not normally done because the lamp power at the excitation wavelength is not always correlated with the total power of the lamp. However, the chromatic noise of the arc might be less severe than its spatial noise. COMPARISON OF THE FRONT SURFACE/REAR SURFACE FLUOROMETER AND CELL SHIFT INSTRUMENTS The front surface/ rear surface fluorometer has some advantages over instruments that use the cell Shift method. The efficiency of the transport and collection of light in this instrument is better that that Of the cell shift instrument, due to the collimation requirement of the latter (2). Also, the use of fiber Optics allows greater flexibility of the optical alignment. For example, the only critical optical axis in the instrument is the lamp/lens/ excitation monochromator axis; any other component of the system may be placed in essentially any position. The use of fiber optics results in good immunity to vibration of the sample cell, which would be important in stopped-flow kinetics measurements and process control applications. Unlike the cell shift instrument, the new instrument is compatible with flowing streams and capable of real-time analysis. The time response of the new instrument for the measurement and display of corrected fluorescence data is limited by the speed at which the calculations can be made, rather than by any mechanical movement. If real time display of corrected fluorescence is not required, the time response of the system is unlimited. Unfortunately, there are also disadvantages to the front surface/rear surface fluorometer. The ability of this instrument to correct for inner filter effects is not as good as the cell-shift instrument. The correction equation for the new system contains an 73 Optical transfer function that was Obtained by curve fitting. Thus, it is not currently known under what conditions these corrections would be expected to work. As explained in Chapter 2, the transfer function is expected to change based on changes in the numerical aperture of the optical system, as well as the refractive index of the solution being measured. Although studies are currently underway to determine the wavelength dependence of the transfer function experimentally, and the efi'ect of the refractive index (6), these are not currently known. As a general instrument for fluorescence spectroscopy, the new design is less versatile than the cell shift instrument. Nephelometry (90’ scattering) is not possible, and turbidirnetric measurements (0’ scattering) would have to be interpreted semiempirically, as the fluorescence was. Measurements that require polarized light are not possible, since the multirnode fibers convert linearly polarized light to elliptically polarized light (7). The cell Shift method provides the operator with two different estimates of the primary and secondary absorbance, whereas this instrument does not. Differences in the absorbance estimates indicate that a compound is adsorbing to the cell walls or that the sample is inhomogeneous (1). Those errors would go undiagnosed with this instrument. SUGGESTIONS FOR FURTHER STUDY Modifications of the Current Instrument In spite of the fact that demand by the communications industry has helped to lower the price of monochromators, an instrument which uses four monochromators is unlikely to be considered practical by many people. Furthermore, the dynamic range was limited on the lower end by stray light in the monochromators, and additional filters were required to achieve reasonable performance. The use of bandpass filters instead of monochromators is therefore encouraged. Interference filters, in combination with shortpass filters (to remove higher orders) would be most useful. The greatest obstacle to 74 the use Of filters would be their wider spectral bandpass, which could cause deviations from Beer’s law at high absorbances. A study of the effectiveness of absorption correction with respect to excitation and emission bandpass would prove useful in any case. If wider bandpasses can be used than were used in these studies, lower detection limits could be achieved, providing the stray light could be reduced. The use of a laser as an excitation source would provide many advantages. Recently, several low cost integrated dye laser systems, which could replace the arc lamp / excitation monochromator combination have been introduced (8). The bandwidth of a laser is quite narrow, and its use would eliminate stray light from the excitation optics. The high power of a laser would lower the detection limit of the instrument, if the high stray light of the emission optics could be lessened. Laser sources can be focused to diffraction limited spot diameters, which makes them ideal for introducing light into a single Optical fiber. Thus the fiber bundles in the current instrument could be replaced with single fiber/beam splitter combinations. This would allow the construction of much smaller flow cells. A 600 um fiber would produce a 2.8 11L flowcell; a 200 um fiber would produce a 0.3 11L flowcell (9). Furthermore, the excitation/emission cone overlap problem described in Chapter 2 would not exist; the transfer function for a single fiber system would probably be monotonic. The disadvantage of a single fiber would be that, since the same fiber would conduct light to the cell and from the cell, reflections at each fiber face would contribute directly to the stray light. Continuous wave (CW) lasers, although more expensive than the pulsed lasers mentioned above, could provide an advantage that would justify the extra expense. With a CW laser as the excitation source, the powerful technique of phase fluorometry (10) could be combined with absorption correction. In phase fluorometry, the excitation source is sinusoidally modulated at a high frequency, typically several tens of megahertz. Most optical modulators that operate at such high frequencies require a highly collimated 75 light source, which is why a laser is necessary. The fluorescence emission intensity will then oscillate at the modulation frequency, but the AC portion of the emission will be phase shifted relative to the excitation beam, and demodulated as well. Either the phase shift or the demodulation factor can be used to calculate the lifetime of the fluorophore. One reason that phase fluorometry is preferred over pulsed lifetime methods is that phase sensitive detection (lock-in amplification) can be used to great advantage. The phase shift 8 between the excitation and emission beams is given by: 6 = tan'1 (or) where w is the modulation frequency and r is the lifetime of the fluorophore. Any contribution to the emission intensity caused by elastic or inelastic scattering appears the same as a contribution from a fluorophore with an infinitely short lifetime. Thus scattered light is never phase shifted relative to the excitation beam. It is possible to modulate the excitation beam at a frequency such that the fluorescence emission is phase shifted by 45° relative to the excitation beam. Phase sensitive detection can then used to discriminate fluorescence from inelastic scattering (11) and Raman scattering (12). The former is especially relevant to this dissertation. It represents a way to minimize the stray light due to the scattered excitation beam, which was the major cause of the large fluorescence blank reported earlier in this chapter. Another advantage phase fluorometry has over pulsed lifetime methods is that it is possible to determine the lifetime of a fluorophore in less than one second by phase fluorometry. This has applications in both qualitative and quantitative analysis. For example, it is possible to determine the lifetime of a fluorescent eluite as it elutes from a column (13). Knowledge of the lifetime aids in the identification of fluorophores with similar spectra but different lifetimes. 76 However, the opportunities that rapid lifetime determinations Offer for quantitative analysis are greater. By dividing the steady-state fluorescence emission intensity by the lifetime of the fluorophore, one obtains a quantity that is independent of the quantum efficiency, and thus independent of the concentration of any species that quenches the fluorophore (14,15). If this Operation were combined with the primary and secondary absorption correction scheme presented in this dissertation, it would be possible, on a one-second time scale, to make fluorescence measurements that are independent of the concentration of any absorbers or quenchers. The final suggestion for modification of the current instrument is the use of an inert, white flowcell. Although a new model might be required, a white flowcell would increase the Optical throughput in somewhat the same way that mirrored cuvettes do when they are used with a conventional fluorometer (16). Due to difficulty in its machining, Teflon would probably be a poor choice; a better one would be ultrahigh molecular weight polyethylene, which is nearly as chemically inert (17). Other Experiments in Absorption-Corrected Fluorescence The instrument described in this dissertation was designed to satisfy the requirements of flowing stream compatibility and real time data acquisition. If the second requirement is relaxed, there are some modifications that could simplify the instrument and correction algorithm. One such modification was suggested earlier: the excitation and transmittance monochromators could be slewed rapidly between the «citation and emission wavelengths, while the front surface fluorescence PMT is protected by a shutter when the «citation monochromator is set to the emission wavelength. The advantage to this change would be the increased accuracy, relative to the present method, with which the secondary absorbance value would be known. The uncertainty in the value Of the secondary absorbance would not depend on the value of 77 the primary absorbance. Furthermore, the second fluorescence channel would not be required. The disadvantages Of this modification would include the sacrifice of real-time performance, the imprecision of the wavelength setting, and the additional need for slewing motors for the two monochromators. Another way to measure the secondary absorbance accurately would be to add a second flow cell, either sequential or parallel, with its own continuum source (ie. a tungsten bulb), monochromator, and detector. All of the techniques mentioned thus far have been based on three separate measurements - either two fluorescence and one absorbance or two absorbance and one fluorescence - because three values are being determined: corrected fluorescence, primary absorbance and secondary absorbance. Recall that in the equation used to correct for inner filter effects, the 2 absorbance values are added together everywhere they appear. In other words, all that is needed is the total absorbance (primary 4- secondary) in order to correct for inner filter effects. If the individual absorbance values are not required, two other configurations can be suggested. In the first, two flow cells of different lengths would be used. If the sum of the absorbances were small, the signal from the long flow cell would be greater than that from the Short flow cell, because the the «citation beam would reach the end of the cell, and the light emitted from there would reach the front Of the cell without attenuation. In the case of high primary or secondary absorbance, or both, the signals from the two cells would be similar, because the solution at the rear of the long cell would contribute little to the measured front surface fluorescence signal. The ratio Of the two signals would be indicative of the total absorbance. Figure 4-3 shows how the ratio of the two signals is expected to change as a function of the total absorbance based on the model used in the «perirnental section. Unlike the ratio Of front / rear fluorescence, the ratio of the long cell fluorescence to the short cell fluorescence can be «pressed analytically and solved «actly. The stray light 78 2.5 - O - p 2.0 E 1.1.! o I 5 Q In Iii g 1.5 - 1:! 1.0 - I . I v I ' I ' I 0.0 1.0 2.0 5.0 4.0 TOTAL ABSORBANCE Figure 4-3. Calculated ratio of front surface fluorescence measurement from 2 cm flow cell to front surface fluorescence measurement from 0.5 cm flow cell. 79 problem would be expected to be less severe with this instrument as there is no solution / glass interface directly Opposite the «citation window. The method proposed above relies on one value of the ratio of the signal measured with two different pathlengths. If the ratio of signals from many cells of different lengths could be obtained, a more reliable corrected fluorescence measurement would result. One way to achieve this would be with a single channel, front surface fluorescence instrument. A variable pathlength would be Obtained by flowing a bubble through the flow cell, or a segment of some liquid which is immiscible with the sample solution. The liquid would also be required to have a much different refractive ind« than the sample. As the bubble or liquid entered the flow cell, the solution to be measured would follow it, resulting in a time-dependent sample pathlength. The interface between the bubble or the liquid and the sample would effectively function as the rear of the cell, because the difference in the refractive indices would guarantee that light would be scattered once it reached that point, and the curved meniscus between the two solutions would mean that most of the radiation would scatter to the walls, rather than back into the solution. Figure 4-4a illustrates how this concept could be implemented. The front surface fluorescence signals expected for two solutions that produce the same signal when they fill the flow cell completely are shown in Figure 4-4b. These signals would be indistinguishable by a single, steady-state measurement. Note that the rise time of the sample with a high total absorbance is less than the rise time Of the sample with low absorbance. It might be possible to derive an equation that relates the rise time of the front surface fluorescence signal to the total absorbance. Of course, with this method, real-time response is sacrificed, and it would only be compatible with sample streams that could be segmented by gas or liquid. The advantage is that the system would require only one source/ detector and would thus be very simple to implement. It is expected that a low noise syringe pump would be required, because with 80 f in IOW out (a) excitation ‘ I ‘—\ bubble bifurcated fiber optic front surface emission bundle 1.0 - g 0.0 - ‘8? E. 0.0 - 0 iii 8 (b) d 0.4 - 9 ii 8 0.2 - 0.0 '- I ' I T l U I U I u I 0.0 0.2 0.4 0.0 0.0 1 .0 LENGTH or SAMPLE SEGMENT (cm) Figure 4-4. Flowing bubble tem: (a) diagram showing time- ' g pathlength, (b) culated front surface fluorescence Si 1 = total absorbance negligible, 2 -= total absorbance of 2.0).21131 81 a peristaltic pump the position of the interface between the sample and the bubble would oscillate. When this project was initiated most researchers using fiber optics for in situ fluorescence measurements used bundles, like the ones that were used in the project. Early workers found that, because of the Often complicated nature Of the sample matrix, the best one could hope to do was to report changes in fluorescence relative to some initial value (18). In many cases it was not clear whether changes were due to changes in the concentration of the fluorophore or changes in the concentration of some absorber in the matrix. Most recent in situ measurements have involved the use of a single fiber for the combined «citation and collection of emission (19,20). The smaller diameter Of a single fiber allows it to reach places inaccessible tO a bundle; a single fiber is also much less expensive than a bundle. In measurements made in vivo, the matrix has been Shown to cause significant inner filter errors. In «perirnents involving the in vivo monitoring Of NADH, an equation was derived that allows absorption correction based on the measurements of reflectance from surrounding tissue (20). With other in situ measurements, however, no surface esdsts for this type of correction (21). The switch to single fibers has made it possible to suggest a way of correcting in situ measurements for inner filter errors. In this case, as in any other, the most reliable method is to make two or more measurements with different pathlengths of «citation and emission. This could be done by forming lenses on the ends of the fibers that are placed in situ (22,23). Two measurements could be made: one with a fiber with short focal length and one with a fiber with along focal length. The latter measurement would be affected more by absorption by the concommitants in the matrix than the former. The ratio of the two signals could be related to the sum of the primary and secondary absorbance as it was with the two pathlength system described above. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 82 CHAPTER 4 REFERENCES Adamsons, K.; Sell, J.E.; Holland, J .F.; Timnick, A. Am Lab. 1984, 16, 16. Ratzlaff, E.H.; Harfmann, R.G.; Crouch, S.R. Anal Chem. 1984, 56, 342. Christmann, D.R. Ph.D. Dissertation, Michigan State University, East Lansing, MI, 1980. Miller, J .N., ed., Standards in Fluorescence Spectrometry, Chapman and Hall: New York, 1981. Model LS-5 Operator’s Manual, Perkin Elmer Corp., Danbury, CT. Victor, M.A.; Ph.D. research in progress, Michigan State University, 1987. Culshaw, B. Optical Fibre Sensing and Signal Processing, Peter Peregrinus, Ltd.: New York, 1984. Laser Science Inc., Cambridge, MA Vurek, G.G. Anal Chem 1982, 54, 840. Lakowicz, J.R. Principles of Fluorescence Spectroscopy, Plenum Press: New York, 1983. Nithipatikom, K.; McGown, L.B. Appl Spectrosc. 1987, 41, 1080. Demas, J.N.; Keller, R.A. Anal Chem 1985, 57, 538. Brushwyler, K.R.; Bright, F.V.; Hieftje, G.M. Abstract #104, FACSS XIV, Detroit, MI, 1987. Hieftje, G.M.; Haugen, G.R. Anal Chim. Acta 1981, 123, 255. Demas, J .N.; Jones, W.M.; Keller, R.A. Anal Chem 1986, 58, 538. Street, K.W.; Singh, A. Anal Lett. 1985, 18, 529. Cadillac Plastic and Chemical Co., Detroit, MI. Chance, B.; Legallais, V.; Sarge, J .; Graham, N. Anal Biochem 1975, 66, 498. Sepaniak, MJ.; Tromberg, B.J., Eastham, J .F. Clin. Chem 1983, 29, 1678. Renault, G.; Raynal, E.; Sinet, M.; Muffat-Joly, M.; Berthier, J.; Cornillault, J .; Pocidalo, J. Adv. Exp. Med Bio., 1986, 121, 229. Chudyk, W.A.; Carrabba, M.M.; Kenny, J.E. Anal Chem 1985, 57, 1237. 83 22. Lee, KA; Barnes, Es. Appl Opt. 1985, 24, 3134. 23. Russo, V.; Righini, G.C.; Sottini, S.; Triganri, S. Appl Opt. 1984, 23, 3277. CHAPTER 5: INTRODUCTION TO DIGITAL FILTERS Most «perirnental scientists are somewhat familiar with the process of filtering analog electrical signals. At the very least, one learns how to set the rise time on a preamplifier or chart recorder to be short enough so as not to distort the signal, and long enough to reject high frequency noise. However, many scientists are not aware that any filtering Operation that can be done by analog circuits can be done by arithmetic operations on a digitized version of the signal. The arithmetic versions, which are known as digital filters, have certain advantages compared to the widely used analog filters. Conversely, many scientists are not aware that some of the arithmetic operations they perform on data, such as smoothing and averaging, are filtering Operations. In many cases, realizing when an arithmetic Operation is a filtering process allows one to achieve better results more quickly than would have been obtained through many hours of trial and error. In this chapter some of the general characteristics of filters are discussed. These concepts will shed some light on the techniques used to analyze digital filters, which are «plained in Chapter 6, and those used to design digital filters, which are «plained in Chapter 7. This chapter is concluded by a survey of the use of digital filters in analytical chemistry. ANALOG FILTERS Digital filters have much in common with analog filters. An analog filter can be characterized by its impulse response, which is the output of the filter when an infinitely narrow pulse, known as the delta function, is applied to its input. From the impulse 85 response, the output of the filter for any input waveform can be deduced. The output is given by the convolution of the impulse response and the input waveform: y(t)=fh(7)x(t-T)d7 (5-1) where x(t) is the input, y(t) is the output, and h(r) is the impulse response of the. filter. The FOurier transform of the impulse response of a filter is given by : X(6I)= :1:(t)e‘""dt where i is (-1)1/2. In this dissertation, i will be used rather than the electrical engineering convention j. X(w) is called the transfer function of the filter, and is a function of frequency. The transfer function is generally more useful than the impulse response for characterizing a filter . The reason for this stems from one of the theorems of Fourier transforms, which states that convolution in the time domain is equivalent to multiplication in the frequency domain (1). In other words, the Fourier transform of the output of a filter is given by the Fourier transform of the input times the transfer function of the filter. Since this multiplication process is much easier to visualize than the convolution process, filters are usually characterized by their frequency domain characteristics. The transfer function is usually converted from its compl« form, A(x) + iB(x), to the magnitude/phase representation, where the magnitude, or gain is given by [A2(x) +Bz(x) ]1/2, and the phase is given by tan“1 [B(x) / A(x)]. In this form, the transfer function is Often called the frequency response of the filter. The magnitude part of the frequency response is the gain the filter has for a sinusoid of a given frequency. The phase part Of the frequency response is the amount by which the phase of a sinusoid of a given frequency is shifted upon passing through the filter. 86 DIGITAL FILTERS All of the characteristics of analog filters listed above also pertain to digital filters, but in a slightly different form. Digital filters are characterized by an impulse response. The effect they have on any signal can be predicted by the discrete convolution function: ya: 2 ckx‘n-k (5'2) kz-aa where the x" values are the input data, the y" values are the output data, and the ck values constitute the impulse response of the filter; that is, the output when the input is: xn=1,n=0 xn=0,n+0 The impulse response of a digital filter is synonymous with the terms convolution fitnction and filter coeflicients. This is because Equation 5-2 is more than a way of predicting what the output of the filter will be; it is also the way the output is Obtained. Unlike analog filtering, which requires the use Of discrete circuit elements such as amplifiers, capacitors, resistors and inductors to achieve the convolution described in Equation 5-1, digital filtering only requires that one do the calculation of Equation 5-2 with digital multipliers and accumulators. There is also a discrete version of the Fourier transform that holds for a finite number Of data points N. It is given by: N-t . n=0 As with analog filters, the Fourier transform of the impulse response of a digital filter gives its transfer function. Unfortunately it is in a discrete form, like the impulse response itself. In Chapter 6 a method for determining the form of the continuous transfer function of a digital filter is given. 87 Note that the discrete Fourier transform only involves a summation from 0 to N-l rather than from -00 to co. This is one Of the many results of the Nyquist theorem, also known as the sampling theorem. This theorem states that a continuous signal which has a Fourier transform that is nonzero above fs / 2 is not accurately represented if digitized and sampled at a frequency of f8. Any components of the signal at frequencies greater than fs / 2 are folded back and interpreted as having a frequency between 0 and f8 / 2. The sampling theorem is difficult to «plain but easy to demonstrate. Figure 5-1a shows what happens when three sinusoids of frequency 10, 30 and 50 Hz are sampled at a rate of 40 Hz. Note that all three sine waves take on the same value at each sampling point. Only the 10 Hz sine wave has been sampled correctly. Figure 5-1b Shows how 30 Hz and 50 Hz fold back into the region between 0 and 20 Hz, and appear indistinguishable from 10 Hz. The frequency fs / 2 is called the Nyquist frequency or folding fiequency. Where analog and digital filter theory differ, the difference is most likely due to some consequence of the sampling theorem. It should be noted that Equation 5-2 is a Simplified version of the general digital filtering formula: yn=k=2mckxn-k "I' E djy‘n-j j=-ea If d]- = 0 for all j, as in Equation 5-2, the filter is called a nonrecursive filter. If one or more d j are not equal to zero, the filter is called a recursive filter. In this chapter and those that follow only nonrecursive filters will be considered. LINEARITY AND TIME INVARIAN CE The relationships between the impulse response of a filter, its Fourier transform, and the Fourier transforms of the input and output signals only hold for linear, time- VIII I e l (b) l 1.. 50 Hz ._ I 50 Hz l I _ T _ I DO 10 Hz 20 Hz Figure 5-1. Illustration of the Nyquist theorem. 89 invariant (LTI) systems. In Order for a system to be an LTI system, it must obey the following three criteria: 1. If inputA results in output A’, and input B results in output B’, inputA + B must result in output A’ + B'. 2. If inputA results in outputA’, input c-A must result in output c-A', where c is a constant. 3. If inputA(t) results in outputA’(t), inputA(t + s) must result in outputA'(t + s), where s is a constant. An analog filter is a LTI system when the input is kept within a range of amplitudes and frequencies. A digital filter is an LTI system if the filter coefficients do not change with time. If they do, the filter is no longer an LTI system, and the results of filtering can vary in unpredictable ways, even when similar signals are filtered. When a filter is not linear or time-invariant, it must be studied by observation of its output when the signal of interest is the input. Chapter 6 contains a study of a particular non-linear, time-variant filter. LTI systems are capable of performing the common filtering Operations of low- pass, high-pass, bandpass and bandstop (notch) filtering. They are also capable of performing differentiation and integration, which are LTI Operations. It was noted earlier that when a sinusoidal signal passes through a filter, a sinusoid of the same frequency emerges, but in general the amplitude and phase of the output is not the same as the input. This is one consequence of the fact that sinusoids are the eigenfunctions Of filter operations. They are also the eigenfunctions Of differentiation: d/dt eiw‘ - be“ where eiw‘ 8 cos wt 4» Sin at (the Euler relation). The output is simply the input multiplied by a constant. Therefore, a differentiator is nothing more than a filter with a transfer function of i0. Since the transfer function rises linearly with frequency, a 90 differentiator is a high-pass filter. This explains the common observation that differentiation degrades the signal to noise (S/N) ratio of most signals; the gain of a differentiator is highest in the region of the spectrum where the S /N ratio is often lowest. Sinusoids are also the eigenfunctions of integration: f eimt 61:48“ W An integrator is a filter with a transfer function of 1 / iw, which makes it a low-pass filter. Integration commonly increases the S/N ratio of a signal. Because an integrator has infinite gain at w = 0, it cannot be implemented by a non-recursive digital filter. ADVANTAGES OF DIGITAL FILTERS Although digital and analog filters have much in common, there are many advantages to the use of digital filters. One of the greatest is that a digital filter need not be causal. A causal filter is one that has an impulse response that is zero for r > 0. In other words, a causal filter does not respond to the delta function until the pulse actually reaches the input terminal. All electronic filters are causal; they are sometimes called realizable filters. In contrast, digital filters are often used on stored, digitized versions of a Signal. In these cases the value of the signal for future time is available; that is, the value of It”) , x“) , etc. can be used in filtering x". Digital filters are usually designed with coefi‘icients that are symmetric about k a 0. These symmetric filters have the important property that the phase shift is zero throughout the entire frequency range. All electronic filters cause a frequency-dependent phase shift, and are therefore more likely to distort a signal. 91 Electronic filters must be implemented with electronic components such as resistors, capacitors, inductors and operational amplifiers. To change the frequency response of an electronic filter, one must change the values of these components. Furthermore, the values of these components can change from the nominal values due to changes in temperature and humidity, aging, and other factors. The frequency response of a digital filter is determined by its coefficients. Changing the frequency response of a digital filter merely requires a change in the coefficients, which can be done by a computer program. Furthermore, the. coefficients, and thus the frequency response, are independent of changes in the environment. Perhaps the greatest advantage digital filters ofl'er the «perirnental scientist is that they operate on a copy of the original signal rather than the signal itself. If a Signal is judged to have been filtered incorrectly by a digital filter it may be filtered again and again until the desired result is achieved. If a signal is filtered incorrectly by an analog filter, the «perirnent must be repeated and the signal regenerated. One should not get the impression that analog electronic filters are no longer required. At the very least an antialiasing filter is needed immediately before the digitizer. An antialiasing filter is a low-pass filter with a cutoff frequency near fs / 2 that prevents the aliasing phenomenon described earlier in this chapter. Furthermore, low- pass filters are sometimes needed to keep the analog signal within the dynamic range of the digitizer. However, most other filtering operations that were once done by analog filters can now be done by digital filters, with better results and more versatility. 92 A BRIEF HISTORY OF CONVOLUTION FILTERS IN ANALYTICAL CHEMISTRY Savitzky-Golay Filters The history of the use of digital filters in analytical chemistry begins with the work Of Savitzky and Golay (2) in 1962. They published several tables of filter coefficients, which now bear their names, that can be used to smooth (low-pass filter) data or calculate the first through fourth derivative of a data set. The coefficients are derived on the assumption that the data can be described by a polynomial Of a certain degree. Savitzky and Golay showed that convolution of the data with the coefficients is equivalent to fitting a polynomial to a range of points, and replacing the center point of the range with the calculated value of the polynomial at that point. Although mathematically equivalent to curve fitting, the convolution process is much simpler and faster. Unfortunately, several Of the coefficients they published are in error; these were later corrected by Steiner et al (3). Further corrections were given by Madden (4), who also provided an equation that allows one to calculate the coefi‘icients of filters with widths greater than 25, which is the limit of the tables published by Savitzky and Golay. Within a few years there were hundreds of reported applications of Savitzky- Golay (SG) filters. Out of concern that the filters were used improperly in some cases, Willson and Edwards (5) and Enke and Nieman (6) thoroughly studied the effects of SG filters on the signal-to-noise (S/N) ratio of Gaussian and Lorentzian peaks, and on the distortion SG filters cause to these peaks. The former is an «cellent tutorial on virtually all aspects Of the handling of digitized data, and is recommended to anyone with an interest in signal processing. The concept of the smoothing ratio was introduced in the latter article. The smoothing ratio is the ratio of the full width at half maximum (FWHM) of the convolution function to the FWHM of the spectral peak. Although the recommendations of the two groups sometimes differ - Enke and Nieman recommend a 93 smoothing ratio of 2 for maximum S/N enhancement, whereas Willson and Edwards recommend a ratio of 1.2 - they are still widely quoted and heeded by those who use 56 filters. Both of these studies were confined to smoothing filters; a similar «amination of derivative filters was done by O’Haver and Begley (7). The studies described above were all empirical, in that S/N ratios and degrees of distortion were determined by the actual use of $6 filters on calculated peak profiles, both with and without added noise. Recent work has focused on the effects that convolution has on the equations that describe these peak profiles, and this work has produced results that are more generally useful (8-12). Gans and Gill (8) derived an «pression that shows that the sum of the squares of the residuals (SSR) measured between a filtered data set and its unfiltered version is composed of a signal distortion term and a noise term. They also showed «perirnentally that a plot of the SSR versus the number Of coefficients in an SG filter has an inflection point at some number of coefficients N,- , where the greatest contributor to the SSR ceases to be the noise and becomes the distortion. They propose that the distortion caused by the filter is minimized, and the noise removal is maximized, by an SG filter Of N,- points. Bromba and Ziegler (9) reported some of the properties «hibited by SG filters. These properties include, for an SG filter of order m, conservation of all polynomials of degree m +1 or less and conservation of the first m +1 moments of the data set. Based on these properties, they questioned the experimental results of Enke and Nieman, who found that the area (zeroth moment) of a Gaussian peak is changed by some SG filters. They reiterated that the factor by which the variance of white noise is reduced by any nonrecursive filter is given by the sum of the squares of the coefficients of that filter regardless of the distribution of the noise, which once again contradicts the «perirnental results of Enke and Nieman. They also stated that, for SG smoothing filters, the noise reduction factor is also given by the square of the zeroth filter coefficient. Gans and Gill 94 (8) showed that the noise reduction factor is given by the square of the nth coefficient for a nth derivative filter. The paper by Bromba and Ziegler also contains the frequency response plots of some SG filters, and the authors point out that the tangency of the gain curve at zero frequency is of order n for an ntll order SG filter. A more comprehensive collection of frequency response plots was published by Betty and Horlick (24). Ziegler (10) compared SG filters to analog RC filters. He recommended that those who practice slow-scan spectrometry replace their analog RC filters with digital SG filters. This paper also contains several useful plots of the calculated attenuation of Gaussian and Lorentzian peaks, and the increase in the linewidths Of these peaks, as a function of the rise time (similar to Enke and Nieman’s smoothing ratio) of SG filters of orders zero through six. Bromba and Ziegler (11) provided a mathematical definition of a typical spectral lineshape. This definition encompasses Gaussian and Lorentzian peaks, as well as many Others. They tabulated the time domain and frequency domain characteristics that a smoothing filter must have in order to conserve the symmetry, positivicity, position, height, width, and range of values (maximum - minimum) of a spectrometric lineshape. They also showed which classes of filters have one or more of these characteristics. Perhaps the most impressive summary of filter performance to date was done by de Blasi et al (12). They showed that the deformation caused by a filter can be estimated at any point in a signal. The higher order terms of a Taylor series «pansion of the signal are calculated; these are multiplied by the corresponding moments of a continuous time version of the impulse response of the filter to yield the deformation. As one might guess from this description of the method, it is complicated and time consuming, but powerful. This paper also includes a plot that gives the amount by which a Cauchy - Lorenz peak is deformed by linear and quadratic SG filters of various widths. 95 Other Symmetric Filters Bromba and Ziegler presented a recursive implementation of 86 filters (13). They later found that these filters were sometimes unstable due tO roundoff errors, a common problem with recursive filters, and derived a new class of filters that can be implemented recursively with little susceptibility to roundoff errors (14). The authors also «tended the filter design to include filters that are capable of resolution enhancement (15). Biermann and Ziegler (16) published plots that allow one to select the parameters of the filter based on the desired height, S/N ratio, and resolution of the output. Gianelli and Altamura showed that SG filters can be replaced by rectangular weighing functions in which the coefficients take on only two nonzero values (17). These filters can be implemented with only two multiplications and n additions to achieve the same effect as an SG filter with a width of n. They also showed, through the analysis method of de Blasi et al, that these simpler filters cause only slightly more distortion than an Optimum filter (18). Kaiser and Reed (19) published equations that allow one to design low-pass and derivative filters. These filters are similar to those described in Chapter 7 of this dissertation. Three types Of filters are described: those with ripples in both the passband and the stopband, those with ripples only in the stopband, and monotone filters with no ripples in the frequency response. For the first two types, the designer specifies the desired cutoff frequency and the maximum allowable magnitude of the ripple. For the third type the designer specifies the cutoff frequency and the width of the transition from the passband to the stopband. In all three cases, the specification determines the number of coefficients in the filter. Marchand and Mamet have discussed the advantages of the monotone filters for some applications (20). 96 Asymmetric Filters In all the work reviewed thus far, the coefi'icients of the filter have been symmetric. A symmetric filter of n points cannot be used on the first or last (n / 2) - 1 points in a data set, so these points are often discarded, unless one is willing to decrease the number of coefiicients in the filter that is used at the extremes (6,21). In an attempt to work around this disadvantage, Proctor and Sherwood (22) and Leach et al (23) designed SG filters in which the polynomial that is fit to a region of the data is used to estimate any point within the region, rather than just the center point. These filters are discussed in greater detail in Chapter 6. CONCLUSIONS From the literature cited above, it is obvious that an understanding of the frequency response of a filter is necessary if the filter is to be applied properly. However, the literature contains no «planation of how one Obtains the frequency response of an arbitrary nonrecursive filter. New filter designs are Often reported without disclosure of their frequency domain characteristics. Furthermore, most of the work on filters in analytical chemistry has involved studies of the effects of S6 filters on spectral lineshapes. There is little mention of how one would design a filter to remove correlated noise or do any of the other operations filters might be called upon to do besides low-pass filtering. The engineering literature contains a vast amount of information about the design and analysis Of digital filters, but much of it is difficult to understand without an electrical engineering background. The n«t two chapters contain an «planation of how to calculate and interpret the frequency domain characteristics of nonrecursive digital filters, and how to design a filter that matches a given frequency domain specification. 109399???) hit-5H Ni"? 13. 14. 15. 16. 17. 18. 19. 20. 21. 97 CHAPTER 5 REFERENCES Bracewell, R. The Fourier TIonsfonn and itsApplications, McGraw-Hill: New York, 1965. Savitsky, A.; Golay, MJ.E. Anal Grern. 1964, 36, 1627. Steiner, J .; Termonia, Y.; Deltour, J. Anal Chem 1972, 44, 1906. Madden, H.H. Anal Chem 1978, 50, 1383. Willson, P.D.; Edwards, T.H. Appl Spectroscop. Rev. 1976, 12, 1. Enke, C.G.; Nieman, T.A. Anal Chem 1976, 48, 705A. O’Haver, T.C.; Begley, T. Anal Chem 1981, 53, 1876. Gans, P.; Gill, J .B. Appl Spectrosc. 1983, 37, 515. Bromba, M.U.A.; Ziegler, H. Anal Chem 1981, 53, 1583. Ziegler, H. Appl Spectrosc. 1981, 58, 687. Bromba, M.UA.; Ziegler, H. Anal Chem 1983, 55, 648. de Blasi, M.; Gianelli, G.; Papoff, P.; Rotunno, T. Ann. Chim (Rome) 1975, 65, 183. Bromba, M.U.A.; Ziegler, H. Anal Chem 1979, 51, 1760. Bromba, M.U.A.; Ziegler, H. Anal Chem 1983, 55, 1299. Bromba, M.U.A.; Ziegler, H. Anal Chem 1984, 56, 2052. Biermann, G; Ziegler, H. Anal Chem 1986, 58, 536. Gianelli, G.; Altamura, 0. Rev. Sci. Instrum 1976, 47, 27. Gianelli, G.; Altamura, 0. Rev. Sci. Instrum 1976, 47, 32. Kaiser, J .F.; Reed, W.A. Rev. Sci. Instrum 1977, 48, 1447. Marchand, P.; Marmet, L Rev. Sci. Instrum 1983, 54, 1034. Khan, A. Anal Chem 1987, 59, 654. Proctor, A.; Sherwood, P.M.A. Anal Chem 1980, 52, 2315. 98 23. Leach, R.A.; Carter, C.A.; Harris, J.M. Anal Chem 1984, 56, 2304. 24. Betty, K.; Horlick, G. Anal Chem 1977, 49, 351. CHAPTER 6 : FREQUENCY RESPONSES OF DIGITAL FILTERS The use of digital filters, including Savitzky-Golay (SG) filters, in analytical chemistry has largely been a matter of trial and error. This is due in part to the inability of the analyst to specify the desired result mathematically, and in part to a lack of understanding of the fundamentals of digital filters. A filter is normally characterized by its frequency response. The frequency response of a filter is a measure of the gain of the filter for a sinusoid of a given frequency, and the degree to which the phase of the sinusoid is shifted. A filter can be characterized further by its noise equivalent bandwidth, which is a measure of the change in the variance of uncorrelated (white) noise that is input to the filter. These concepts apply to digital filters as well as to analog filters. In this chapter, techniques for the analysis of digital filters are «plained; appropriate applications are illustrated by some examples from research at Michigan State University. As «plained in Chapter 5, convolution with a time-invariant convolution function is a linear, time-invariant (LTI) operation. The equation for this Operation is: 00 yn=k2mckxn-k (6-1) where the ck values constitute the convolution function. In order to determine the frequency response of a filter, we need to perform this convolution on the function xn = em”, which is the eigenfunction of all LTI systems. The result is: y“: 2 ch et'o (ft-k) - = 810112 eke-50k k=- k=-°° which demonstrates that cm" is indeed the eigenfunction of a digital filter. The transfer function of the filter is obtained by dividing the output by the input eiw", and is given by: 99 100 2 eke-10k = 2 ck(coskt.I-isinko) : A+Bi (6-2) kz—ss 53"“ This is the polar form of the transfer function. It is converted to the frequency response through the following transformation: ' k Gain: r——A2+B2 = V (kg-“chem k0) + (kg-”chain or) kz-ae Phase=arctan(B/A)=arctan (——.-, i 0,, sin lco ) (6'3) cpcos km in... When a filter has symmetric coefficients (ck - -ck), the sine terms in each summation above equal zero, since the sine is an Odd function. In these cases, which include the derivative SG filters of even order (smoothing, 2nd derivative, 4th derivative, etc.), the equations are: Gain: .4 = 2 01,008 km b..-” (6-4) Phase=11rct.an(0/A)=0o The fact that the symmetric filters cause zero phase Shift throughout the spectrum is important. It results in the characteristic that symmetric filters will not cause a shift in the location of the maximum value of a Single symmetric peak. When a filter has antisymmetric coefficients (ck -= -c_k), the cosine terms in equation 6-3 sum to equal zero, since the cosine is an even function. The equations for the frequency response are then: 101 Gain=B=2 ck sin kw k=-¢5 , (6-5) Phase=arctan(B/0)=90 These equations apply to odd-order-derivative Savitzky-Golay filters (1St derivative, 3"d derivative, etc.). The frequency responses of filters that calculate the first or higher derivatives can be normalized to make them easier to interpret. The gain calculated by equation 6-4 or 6-5 is divided by the true value of the derivative, which for the nth derivative is (iw)". Where the normalized frequency response is one, the derivative is accurate. Where it is greater than one or less than one, the derivative is overestimated or underestimated, respectively. Note that the normalized frequency response is undefined at w=0; it is normally set to one. The other way a digital filter can be characterized is by its noise equivalent bandwidth (NEB). The variance of a signal is related to its power spectral density P(w) by: 02: gluon.) (6.6) where the power spectral density (PSD) is the square of the transfer function. Furthermore, the power spectral density of the output of the filter P'(w) is related to the power spectral density of the input P(w), by: P.(or=|ri(n)|2 10(6) (6'7) where H ( or) is the transfer function of the filter. Combining equations 6-6 and 6-7 we obtain: 102 062 = zirtfP (0M0: 21'le(61)|2 PM do) (6-8) For white noise the power spectral density is constant, so it can be placed outside the integral. Furthermore, we can substitute equation 6-2 into equation 6-8 as the transfer function to get: —:°2—- 2117 f 0,,(008 Ice-i sinktd) k:-c The square of the transfer function, which is what remains inside the integral, is composed of three kinds of terms: cos mw cos no, sin mw sin no and cos no) sin no. The only terms that evaluate to a nonzero integral over the interval (1r, -1r) are the terms of the form cos2 kw and sin2 kw. Thus, ~11 k=-oo and 2 co 0 2 T00 = 2 Cr: (6-9) Therefore, the variance of white noise that is filtered by a digital filter is reduced by a factor equal to the sum of the squares of the coefficients of the filter. These two ways of characterizing a filter are in a sense complementary. The frequency response of a filter helps one to determine if its use will severely distort the underlying waveform of a signal, be it a spectral peak, an exponential decay, or any other waveform. If the gain of the filter is low in a region of the spectrum where the power spectral density of the waveform is high, distortion is likely to result. However, the distortion of a waveform does not always destroy the information that the waveform 103 contains. For «ample, it was stated earlier that a symmetric filter, no matter what its frequency response, will not shift the location of the center of a Single symmetric peak. The frequency response also indicates if the filter will be able to attenuate correlated noise, which is usually confined to a narrow region of the spectrum. However, the frequency response does not completely characterize a filter. For «ample, if one wishes to avoid distorting a waveform, one could filter it with a digital filter that has a gain of one at all frequencies. While this would indeed avoid distortion, it would not produce the desired efl'ect of filtering. This filter, known as an all-pass filter, has coefficients given by: Ck = 1, k = 0 ck -= 0, k 11 0 which results in a noise amplification of 1.0 as calculated via equation 6-9. Thus, this filter does not distort the waveform, nor does it remove noise; in fact, it does nothing, as one would «pect. Although this an extreme «ample, it illustrates that tradeoff between noise reduction and distortion that must be faced by anyone who uses a filter. Greater noise reduction (smaller NEB) can only be achieved through reduction of the integral of the transfer function, which is achieved by lowering the cutoff frequency of the filter, which results in greater distortion of the underlying waveform. EXAMPLES Savitzky-Golay (SG) Filters Savitzky-Golay filters are symmetric or antisymmetric, depending on the order of the derivative being calculated, so that equations 64 (even order derivative) and 6-5 (odd order derivative) can be used to calculate the frequency response. Aside from the order 104 of the derivative, there are three other variables that can be specified to produce the desired frequency response: the order of the polynomial that is used to fit the data (linear, quadratic, cubic, etc.), the number of coefficients in the filter, and the number of passes Of the data through the filter. The effect that each of these variables has on the frequency response will be «amined below. Figure 6-1 shows the frequency responses of linear, quadratic and quartic 1 1 point SG filters. It is Obvious that increasing the order of the polynomial causes the cutoff frequency to move to higher frequencies. Higher order filters also «hibit greater tangency at w=0. Figure 6-2 Shows the frequency response of 7, 11 and 13 point quadratic filters. For a given order, an increase in the number of points in the filter causes the cutoff frequency to move to a lower frequency, and increases the number of ripples in the stopband. Figure 6-3 Shows the frequency responses Of 1, 2, and 3 passes of an 11 point quadratic 86 filter. With each pass of the filter the cutoff frequency decreases, the transition band widens, and the stopband ripples decrease in magnitude. These frequency response plots are readily calculated by taking advantage of the equivalence of convolution in the time domain and multiplication in the frequency domain. If a Single pass of a filter has a frequency response of H(w), then the effect of 11 passes of the filter has a frequency response of [H(w)]". Figure 6-4 shows the frequency response of an 11 point linear differentiator, and the normalized frequency response of the same filter. The advantage of the normalized representation is easily seen. Digital Up/Down Integration It is shown in Chapter 5 that a differentiator is a high pass filter. When differentiation is done by analog circuits, the output is often extremely noisy. For this reason, an electronic circuit was designed that incorporates the noise-rejection 105 1.0- 0.0 0.5 FREQUENCY Figure 6-1. Frequency responses Of.1 1 point Savitzky-Golay filters: (a) linear, ' (b) quadratlc, (c) quartic. 1.0- 106 (a) .1" Figure 6-2. 0.5 FREQUENCY Frequen re uses of quadratic Savitzky-Golay filters. (a) 7 point.(b (139 9059139 (9)11 point- - 107 1.0 - (a) GAIN (b 0.0- (C) ‘ A 4. ~ v 0.0 0.5 FREQUENCY Figure 6—3. Freqmen responses Of multi le asses of a quadratic 11 point Saviltzkyzolay filter: (a) 1 pas}; (g) 2 passes, (c) 3 passes. 108 1.0-1 0.0 0.5 FREQUENCY Figure 64. Frequency response of 11 point quadratic first derivative Savitzky- Golay filter. 109 characteristics of integration to estimate the derivative (1). The method is known as up / down integration. The principle of the method is that the derivative of a signal may be estimated by the slope of a line drawn between two points equidistant from the point at which the derivative is desired. If the actual measured value of the Signal at these two points is used, the slope of the line is bound to be influenced by random errors. A less noisy estimate of the signal at each endpoint of the line can be made if the signal is integrated over some interval surrounding this point, and the integral is divided by the width of the interval. The algorithm is as follows: 1. Integrate the signal at time tl for duration At and divide by At . 2. Integrate the Signal at time t2 for duration At and divide by At . 3. Subtract the integral at time t1 from the integral at time t2. 4. Divide by (t1 - t2) to estimate the derivative. Since the up/down method has proven itself in several inplementations for stopped-flow kinetics (1,10,11) this seemed like a natural place to start as a way to calculate the derivative for digitized stopped-flow data (2). In these experiments, the integral was calculated via Simpson’s rule, yn = y,” + 0.5(xn +1 + x"). Although this is a recursive formula for the derivative, it can be recast into a nonrecursive form, because yo = 0, and the duration of the integral is finite. A simple «ample will demonstrate this. Let us suppose that the point at which we derive the integral is at the center of a set 9 points, with the center point known as x0. Simpson’s rule gives the area of the region beyond the center point as : 110 area2=[ 0.5(x1-1-xo) + 0.5(x2+x1) + 0.5(x3-1-x2) + 0.5(x4+x3)] Ax :-=(051c0+1t1 +x2+1t3 + 0.51:4)Ax The area of the region before the center point is given by: area1-[ 0.5(x_1+x0) + 0.5(x_2+x_1) + 0.5(x_3+x_7) + 0.5(x_4+x_3) ] Ax =(0.5!0 +x_l +x_2 +x_3 + 0&4)” The difference in the two areas is: areaz-area1=( 0.5x4 + x3 + x2 + x1 -x_1 -x_2 -x_3 - 0.51:4) This quantity must be divided by 4Ax to convert the difference in areas to a difference in mean values, then again by 4A1: to convert the difference in mean values to the dam of the line connecting those mean values. The final result for the estimate of the derivative is x’o=( 0.34 ‘I‘ X3 "I' x2 '1" x1 'x_1 “x.2‘x-3' 0.5x_4)/16Ax Note that previous values of the integral y" are not represented in this equation; thus the recursive formula for the up/down integration has been recast in a non-recursive form, which can be represented as follows: xn=( l )ickxn-h m2 Ax kz—aa where ck=0.5 for k=-m, ck-l fory-m4 " .. L I 00 o. o oo o. o. D .0. o e o o e e a .- a fin! - O O o o a O o o n o O o o o e o e o I- n u o. e. on u o. o. s. -. - . 4 o e o. ‘0 - . . .0 O 0'... .0” 0’... a”. ’0’... . a a"... 00° '00. '00. on ’0' '00. as u. v 3, .. .— E '. l < e - 0. 0'... C”. d 4 b I I l J I I I l I I I 1 I r I I Figure 6-10. (arbitrary units) Effect of sliding window filter on noise-free sine wave. 122 maximum are shifted in time (or equivalently, phase shifted) relative to the first minimum and last maximum in the original sine wave. The center of the data set still has the same frequency and phase as the original sine wave, and has an amplitude that is ca. 0.16 times the original amplitude, which is expected, since (0.7)2 a 0.16. In the data set that has been filtered 20 times, the edge effects extend even further into the data set. In the center of the data set, the sine wave has nearly disappeared. It is expected to have an amplitude of (0.7)20 which is ca. 8 x 10 ‘4. It is apparent that the data deviate from this small sine wave as far as 25 points from either end. Most important is the fact that the value at the ends have changed little from the values they took after one pass of the filter. Figure 6-11 shows the effect that the sliding windows have on pure noise. Again the 11 point quadratic filter was used. The points are drawn from a Gaussian distribution of random noise with a mean of zero and unit variance (9). The line with the large oscillations results from smoothing the data 20 times, while the smoother line results from filtering the data 100 times. Since the mean of the noise is zero, an infinite number of passes of the symmetric filter is expected to reduce the data to a straight line with a slope of zero. Near the center of the data set that is clearly happening, but once again, because of the large gains associated with the initial and final point filters, the end values do not change much from the unfiltered values. Since the data are uncorrelated, the phase shift of the asymmetric filters does not contribute to the edge effects in this data set. Figure 6—12 shows the performance of 2 different 21 point quadratic first derivative filters: initial and symmetric. Figure 6-l2a is a plot of absorbance versus time obtained with an automated stopped-flow spectrophotometer (12). Figure 6-12b shows the derivative calculated by the two filters mentioned above. Since the symmetric filter Amplitude 1% _2 -- q r I n I l l l -3. I I ' l ' I ' 1 O 20 4O 60. 80. 100 Point Number Figure 6-11. Effect of sliding window filter on noise. 124 0 5 ' I 1 ‘ I 0.4 -- A D ., < o 0.3 - Raw Data 0 C O f 0.2 u u 0 a d .D < 0.1 - - (a) o 4 : : : . 1|- (b) 4 0.3 -" . . u ’3 lmtlal Point O a 1 \ D 5 o 2 -- - ‘0 '0 \ < .0 0.1 -- u Symmetric 1’ - . J . I I 0.0 v ' I I I I I o. 1 2. 3. 4. Time (see) Figure 6-12. Effect of initial slope filter on stopped-flow data: (a) stopped-flow transient, (b) calculated derivative. - 125 cannot calculate the derivative at the first 10 points in the data set, and the initial point filter cannot calculate the derivative at the last 20 points, these points are not shown. The variance of the derivative calculated by the initial point filter is obviously much greater than the variance of the derivative calculated by the symmetric filter. Furthermore, the phase shift introduced by the initial point filter can be seen in Figure 6-12b. Several minima and maxima in the lower peak are also present in the upper peak, but they occur earlier in the upper peak due to the positive phase shift of the initial point filter. Of course, it is not suggested that an initial point filter be used when the symmetric filter can be used. This example merely shows that, when an initial point filter is used, one should not assume that the variance of the resulting estimate is the same as the variance of an estimate produced by the symmetric filter. CONCLUSIONS The initial point filters provide a good example of why it is necessary that one have a knowledge of the frequency response of a digital filter one intends to use. Many people probably assumed, as we did at first, that because these filters are derived using the same assumptions used to derive the symmetric SG filters, they will perform the same. Analysis of the initial point filters shows that under some conditions they do not act like filters at all. It is hoped that this chapter has demonstrated the ease with which frequency response plots can be generated, and how useful they can be to users of nonrecursive filters. i" PWNP‘S"? 10. 11. 12. 126 CHAPTER 6 REFERENCES Cordos, E.M.; Crouch, S.R.; Malmstadt, H.V. Anal. Chem 1968, 40, 1812. Wentzell, P.D.; Crouch, S.R. Anal. Chem 1986, 58, 2855. Hamming, R.W. Digital Filters, 2nd ed., Prentice-Hall: Englewood Cliffs, NJ, 1983. Nevins, TA; Pardue, H.L.Anal. Chem 1984, 56, 2249. Proctor, A.; Sherwood, P.M.A. Anal. Chem 1980, 52, 2315. Leach, R.A.; Carter, 0A.; Harris, J.M.Anal. Chem. 1984, 56, 2304. Baedeker, PA. Anal Chem 1985, 57, 1477. Wentzell, P.D.; Doherty, T.P.; Crouch, S.R. Anal. Chem 1987, 59, 367. Wentzell, P.D.; Crouch, S.R. J. Chem Ed, submitted for publication. Ingle, J.D., Jr.; Crouch, 8.12. Anal. Chem 1972, 42, 1055. Holler, FJ.; private communication. Crouch, S.R.; Holler, Fl; Notz, P.K.; Beckwith, P.M. Appl. Spectrosc. Rev. 1977, 13, 165. CHAPTER 7 : DESIGN OF DIGITAL FILTERS The preceding chapter detailed the procedure used to determine the frequency response of any nonrecursive digital filter. In many cases the Savitzky-Golay (SG) filters, which were the focal point of the previous chapter, are adequate for any low-pass filtering or differentiation that may be required. At other times, a filter is required that more accurately mimics the ideal low-pass filter; that is, one that has a gain of exactly one for a range of frequencies, and a gain of zero for the remainder. In this chapter, techniques for the design of a filter with a given frequency response are demonstrated, and several examples of their use are shown. These filters are sometimes known as windowed ideal low-pass (WILP) filters. Much of the material in this chapter is based on two excellent and readable textbooks on the subject (1,2). THEORY It has already been stated that, because of the sampling process, a digital filter is characterized by a transfer function that is periodic, with a period equal to the sampling period. Because the transfer function is periodic and continuous, it has a Fourier expansion. In other words, the continuous transfer function can be described by a discrete sum of weighted sines and cosines. This Fourier expansion, which is specified in the time domain, is equivalent to the frequency domain transfer function in the sense that convolution in the time domain is equivalent to multiplication in the frequency domain. Therefore, the Fourier expansion is the set of filter coefficients that must be convolved with the data to obtain the desired filtering. The Fourier expansion of a function H ( w) is given by: 127 128 11(0): g9- +k21ak cos ka+bksin lea where the “k and bk are the time domain coefficients of the expansion. We wish to turn this equation around; namely, given the frequency response, H(w) we want the values of “k and bk' They are determined by the following set of equations: “In: 711' f H(0)cos lea) do; (7-2) bk :71!” f H(w)sinlcm do; (7-3) The results are converted to functions of the sampling frequency through the relation, to g 21f, EXAMPLES Low-pass Filters If a low-pass filter is required, the frequency specification is: H(w) = l, wc< w < a)c H(w) = 0, w > wc,w < -wc where we is the cutoff frequency of the filter. Note that this is an even function, while the sine is odd. The product of the two, which is the function to be integrated in equation 7-3 to obtain the coefficients b k’ is therefore odd, and its integral over a region symmetric about u = 0 is zero. Thus, the value of all the coefficients bk is zero. To evaluate the coefficients of a k we need to solve the equation: ,: — “-1. 129 ”c n . ak=7lffH(a)coslcm do = 711' coslca d0 -II '"c =1|:.1_sinlcm]me = 23inlcan = 28inlcfc 7r k _0 Fit- ° 75—1? o 7-4 (0 ”c _ 1 .. 2 _ aa-fl'fdm "' Tn - 4f¢ "we If one wishes, for example, a filter that passes all frequencies lower than one tenth of the sampling frequency, then fc - 0.1, and the coefficients are given by: Ck = 0.4, k = 0 ck = 2-(sin 0.21rk)/1rk, k If 0 Two points must be made here. First, the equations above hold for all k. Remember that, as shown in equation 7-1, the Fourier expansion of a frequency response involves a summation to infinity over k. If one wishes to design a filter that has a frequency response that matches the specification exactly, the filter must have an infinite number of coefficients. Obviously, the use of such a filter would require an infinite number of multiplications and additions, as well as the loss of an infinite number of data ponits at each end of the data set. Therefore, one is forced to evaluate the coefficients only out to some practical value of k (and -k). The minimum number of coefficients one will use in the actual filter depends on the cutoff frequency and the desired attenuation rate in the transition band. The maximum depends on the rate at which output from the filter is desired and the number of data points that one can afford to lose at each end, since both depend linearly on the number of coefi'icients in the filter. The second important point is that this design equation does not guarantee that the output of the filter is equal to the input when the input is constant. Normally one 130 wishes the output and input to be equal under those conditions, since it guarantees that the DC component of a waveform will not be distorted. Therefore, to avoid DC distortion, after the coefficients are calculated, they should be normalized so that: 2a=1 Figure 7-1 shows an 11 point filter with cutoff frequencies at 0.1, 0.2, and 0.3 times the sampling frequency. In all three curves, the gain starts at one, but soon begins to oscillate. Unlike the SG filters, these filters contain ripples in both the passband and the stopband. These ripples occur because the transfer function contains a discontinuity, where the gain goes from one to zero. In most cases such a sharp cutoff is not needed. A better choice would be to have a transition region of finite duration, and have the transfer function decrease linearly from one to zero in that region. While it would be possible to find the Fourier expansion of such a frequency specification with equation 7-3, there is an easier way to achieve the same result. If the discontinuous transfer function were convolved with a narrow rectangular function in the frequency domain, the result would be exactly what we want: a transfer function with a transition band where the gain decreases linearly. Because of the equivalence of convolution in the frequency domain and multiplication in the time domain, this same result can be obtained if the filter coefficients are multiplied by the Fourier expansion of the small rectangular function, which is given by: ak = N~sin(1rk / N)/1rk do = 1 This process is known as windowing the coefficients. There are many windowing prowdures, and all of them offer a compromise between the width and the shape of the transition region in the transfer function, and the height of the ripples in the transfer I 'i- 131 1.0 - E (b) (0 § l 0.0 0.5 FREQUENCY Figure 7-1. Frequency responses of unwindowed 11 point filters: (a fc = 0.1 times sampling frequen , ) f = 0.2 tnmes sampling equency, (c) fe = 0.3 times sampling eqniency. 132 function (1,3). Figure 7-2 shows the windowed versions of the same filters shown earlier in Figure 7-1. Each of the filters now has a much wider transition region (less sharp cutoff) than the filters in Figure 7-1, but the ripples in both the passband and the stopband have been geatly reduced. Because of the geater accuracy of the windowed filters near DC, they would be preferred in most applications. Differentiators As was stated in Chapter 5, a perfect differentiator behaves as a high pass filter. The signals that are analyzed by chemists often have power spectral densities (PSD) that are concentrated at lower frequencies. The PSD at high frequencies is due mostly to the noise in the signal. Therefore, a perfect differentiator is seldom called for; such a filter has its highest gain for frequencies where there is little signal power density. A useful differentiator must have a cutoff frequency beyond which its gain is close to zero. The derivative of the eigenfunction of an LTT system is given by: d/dn = cw" = imam and the transfer function (output/input) is thus: imam NW = in» This is easily extended to any derivative; the transfer function of the N‘11 derivative is (11.2)”. This transfer function can be substituted into Equations 7-2 and 7-3 to obtain the filter coefficients. If the derivative is odd (1“, 3", etc.), the transfer function is odd. The product of the transfer function and the cosine, which is even, is then odd. Thus, a k is zero for all k. If the derivative is even, (2“, 4th, etc), bk is zero for all 1:. For a symmetric filter it is always true that either all the “I; are zero or all the bi: are zero, and the coefficients of the filter are the ones that are not equal to zero. 133 1.0 " (a) (b) (C) 2 5 0.0 - 4‘.— A—“:. e.._ 0.0 0.5 FREQUENCY Figure ‘7-2. Frequen? responses of the windowed versions of the filters shown in Figure -1 134 The results are, for the first derivative: _ gi ( sin 21Tlctc _ cos 21mg, ) blc' 11 k2 k 7-5 50: 0 and for the second derivative: _ 2 2 at: 131!!— sin 21:ka - 47% cos 21w, ’63 ’52 7-6 “0: 0 These filters are normally windowed in the same way the low-pass filters are. However, there is a different set of normalization criteria associated with derivative filters. We are no longer concerned about the effect of constant input, since the derivative of a constant is zero. Instead, we are concerned about what is obtained when a polynomial of degee m is differentiated. An accurate differentiator would give the result that if xn = n'", a'"/(dn)"' x" = m!. Therefore, derivative filter coefficients are normalized so that z (ck k”) = M! For example, a first derivative filter requires that Z (ckk) =1 Figure 7-3 shows the frequency response of a windowed 21 point first derivative filter with cutoff frequency of 0.1 times the sampling frequency. The curve with a gain of one at w=0 is the normalized frequency response, as described in Chapter 6. Aside from the lack of ripples in the stopband, it looks very much like the differentiator shown in Figure 6-4. 135 1.0 '- normalized i 0.0 l “I. 0.0 0.5 FREQUENCY Figure 7-3. fregnllency response of windowed 21 point derivative filter with c: 0 O 136 WHY DESIGN FILTERS? We have seen that $6 filters are characterized by large ripples in the stopband, whereas windowed ideal low-pass (WILP) filters have much smaller ripples that are distributed throughout the transfer function. This difference endows the SG filters with greater accuracy near DC, which means they would cause less distortion that the WILP filters, although the extra distortion caused by WILP filters can be made arbitrarily small. However, the presence of the ripples in the stopband could cause the SG filters to perform worse than the WILP filters on data in which the noise is not white, but concentrated in certain areas of the spectrum. In this case the noise equivalent bandwidth (NEB) of the filter is irrelevant; what is important is the gain of the filter at those frequencies where the noise power is geat. One cannot treat 86 filters as general low-pass filters, because one cannot be sure that the gain of the filter is approximately zero for the high frequencies. Because WILP filters do not have large ripples anywhere in the spectrum, they are easily converted to other types of filters, including high-pass, bandpass and bandstop (notch) filters. While these filters could be designed by calculating the Fourier expansion of the desired frequency response, it is simple to obtain them from converted low-pass filters. For example, high-pass filters are useful for analyzing a rapidly changing signal that is superimposed on a slowly changing backgound (4). To be converted to a high- pass filter, a low-pass filter must be changed so that the gain is one everywhere it was zero, and zero everywhere it was one. This can be done by subtracting the low-pass transfer function from a transfer function that is equal to one at all frequencies. Recall that this is an all-pass filter, which has the filter coefficients c0 8 1. One of the theories of Fourier transforms states that the Fourier expansion of the sum of two functions is given by the sum of the Fourier expansions of the individual functions. All one need do is reverse the sign of each coefficient except the central one and subtract the central 137 coefficient from one to obtain a high-pass filter with the same cutoff frequency as the original low-pass filter. Bandpass and notch filters can also be designed in this way, although it is a trial and error process. A notch filter will serve as an example. Here, the gain is required to be zero for all frequencies located within a distance Af of fc , the center frequency of the stopband. A filter having approximately this frequency can be made by the addition of a low-pass filter with a cutoff frequency of fc - Af and a high-pass filter with a cutoff frequency of j"c + Af . The part that makes this a trial and error process is that this design algorithm does not guarantee what the value of the gain of the filter will be at fc. EXAMPLES Correlated Noise Removal Figure 7-4a shows a peak that was obtained with a flow injection analysis (FIA) (8). The height of the peak is a measure of the absorbance of the solution, which consisted of a plug of dye injected into the carrier stream. These data were collected as part of an experiment to determine the effect of temperature on dispersion in FIA (5). The sinusoidal component in the peak is pump noise, although this was not immediately apparent. The sampling frequency for these experiments was 1.43 Hz, and the sinusoid has a period of approximately 16 points, which gives a frequency of 1.43 / 16 = 0.09 Hz for the sinusoid. This frequency is much to low to be ascribed to the rollers in the pump, which for this experiment revolved at approximately 1.5 Hz. However, due to aliasing, a signal at 1.52 Hz actually shows up at 0.09 Hz. Since the sinusoid is composed of approximately 16 points per cycle, it has a frequency that is approximately 1/ 16 (0.0625) times the sampling frequency. To remove this noise with minimum distortion of the FIA peak, a bandstop filter with fc = 0.0625 is required. 138 (a) ABSORBANCE -) (b) TIME -> Figure 7-4. gleakcgbtained on flow injection analyzer: (a) unfiltered, (b) ter ‘ 139 In this case a compromise must be made because the data set consists of just 100 points. Although the peak is near the center of the data set, and an estimate of its height is all that is required, one cannot afford to lose more that about 15 points at each end. Figure 7-5 shows the frequency response of the filter that was designed to remove the correlated noise, and the high-pass and low-pass filters that were combined to produce the bandstop filter. Note that the gain of the filter is not zero at fc. To obtain a lower gain at fc, one would have to use more filter coefficients, which would cause the loss of more data at each end of the data set, or specify a lower fc for the low-pass filter and a higher fc for the high-pass filter, which would cause the bandstop region to be wider and the peak to be more distorted. The filter that was used here was found to be a good compromise between distortion and correlated noise rejection. Figure 7-4b shows the FIA peak after it has been filtered by the bandstop filter. The peak height has been decreased by about 4 times the amplitude of the correlated noise. The ratio of FWHM to the peak height has increased by about 25%. Both of these factors indicate that the peak has been distorted somewhat. However, examination of the baseline in Figure 7-4a and 7-4b shows that the amplitude of the correlated noise has been decreased dramatically. The distortion caused by the bandstop filter would not be of geat concern in typical FIA applications. The shape of an FIA peak is largely independent of the concentration of the analyte in a typical assay where the reaction goes to completion. Since the peaks produced by different analyte concentrations are related by a scaling factor, and the bandstop filter is an LTT system, the outputs of the filter are also related by a scaling factor. However, the heights of the filtered versions would be eXpected to be more linear with concentration than the unfiltered versions, since the correlated noise source has random phase relative to the peak maximum. 140 use see: 3 £83 9.3.52 3V cue—m $8.32 A3 .«e 8809.2 .653er .3 25E .6 a q u 3 Jo >ozm30mmn. a... W n I a... g 9 I, 11' 0.. w a N t I...- a E 0.. 141 The bandstop filter is expected to perform better than baseline subtraction because the filter works well independently of the phase of the correlated noise and its amplitude. Its use requires no human invention, and a new filter could be designed automatically to cope with changes in either the pump speed or sampling rate. Decimation There are other cases when 86 filters are inadequate due to the ripples they possess in the stopband. In this section it will be shown that changing the number of data points in a set requires the use of filters that are flat throughout the frequency spectrum (6). Decreasing the size of a data set by retaining only a fraction of the original signal is known in the field of electrical engineering as decimation. Although this term literally means retention of every tenth point in a data set, retention of every N‘11 point is called decimation by a factor of N. Decimation is employed when the signal is sampled too quickly, producing a quantity of data that is unmanageable. Decimation by a factor of N is exactly analogous to having sampled the original data at l/Nth the original sampling frequency. The process of decimation produces a new Nyquist frequency that is 1/N times the original Nyquist frequency. Any components of the signal or noise that existed at frequencies geater than the new Nyquist frequency will be folded back into the spectrum. The process is illustrated in Figure 7-6 for decimation by a factor of two. Note that the process of throwing out every second point causes the frequency axis to be cut in half, and the power spectral density of those components in the upper half of the spectrum folds back and adds to the power spectral density (PSD) of the components in the lower half of the spectrum. In many cases this process causes only insignificant 142 remove every noise folds other point back —” _/ filter remove every other point # amplitude l l frequency Figure 7-6. Decimation by a factor of two. ‘ 143 changes in the spectrum and appearance of the data. However, there may be cases where the original sigial had significant power spectral density at the highest frequencies in the spectrum. This is more likely as the degee of decimation increases, because then the fraction of the old spectrum that folds into the new spectrum increases. Even if significant amounts of noise become aliased to lower frequencies, subsequent filtering can often be used to remove it. However, if the frequency of the noise is such that it folds back into a region of the spectrum where the underlying waveform has high power spectral density, any attempts to remove that noise will result in distortion of the waveform. For these reasons it is essential that the signal be filtered mm decimation occurs. Figure 7-6 demonstrates that, if the proper filter is used, the aliasing of noise and signal components can be prevented. Just as the last step before sampling should be an analog antialiasing filter, the last step before decimation should be a digital antialiasing filter. Figure 7-7a shows a portion of the mass spectrum of Xe obtained with the CVC Model 2000 time-of-flight (TOF) mass spectrometer. These data were collected as part of a study of the effects of suboptimal time lag focusing on TOF data (9). On this system, sampling is performed by a boxcar integator. The output of the boxcar integator contains a correlated noise component that is of unknown origin, but is believed to originate within the spectrometer. This data set consists of 4064 points, and represents a typical oversampling situation: the signal was sampled rapidly so that nothing would be missed, but the high sampling rate resulted in a nearly unmanageable amount of data. If only every 10th point is retained, the spectrum in Figure 7-7b results. Note the presence of a sinusoidal component. This component does not appear to be in the original spectrum. Closer examination reveals that this component is the aliased version of a high frequency sinusoid that is the cause of the 'fuzziness" of the signal in Figure 144 «ascensaeiah on” u C o I ‘ Q Ibvffifihifivfiufi.‘ .. C ..'\W ) a S. b _ I .2 v . . . o...\.‘ s. .. I.‘. JK \A s. on N a . I § v . I O. . C o a o )- f. n. .n. r. . ¢ ~ 0. 0,, \\L V o .. ‘ ~ 1’! «a J . n .. .b 5&3. ‘ T CLwZMFZ. FLIGHT TIME -) of xenon: (a) original spectrum, (b) tering. t mass 10, no Time—of-fli decimated Figure 7-7. 145 7-7a. The sinusoid now has a period that is approximately equal to the FWHM of the peaks. If a bandstop filter were used to remove the sinusoid after the decimation, it would significantly distort the shapes of the two peaks. Since we plan to decimate the data set by a factor of 10, it should be filtered first by a low-pass filter with a cutoff frequency of one tenth the Nyquist frequency, or 0.05 times the sampling frequency. This is a large data set, with large stretches of baseline at each end, so a filter with a large number of coefficients may be used in order to obtain a sharp cutofl'. Figure 7~8a shows the frequency response of the 41 point filter that was used, and Figure 7-8b shows the result of filtering followed by decimation. The spectrum in Figure 7-8b looks to be much closer to the average of the fuzzy spectrum in Figure 7-7a than does the spectrum in Figure 7-7b. Furthermore, the FWHM of the peaks'in Figure 7-8b is the same as in Figure 7-7a, because the peaks’ power spectral densities were small in the upper 90% of the frequency range. Interpolation This same analysis of the effect of changes in the sampling rate on the PSD of a signal can be used to provide a multipurpose method for interpolation. Interpolation can be viewed as increasing the effective sampling rate of a system. We interpolate a data set by a factor of N by adding N - 1 new data points between the original sampled data points. Interpolation schemes difi'er by how the magnitudes of the new data are determined. Consider the effect of placing a data point with a value of zero between each point of original data. Since the data still represent the same time duration, but there are twice as many data points, the sampling rate has been increased by a factor of 2. The PSD of the signal now extends to a Nyquist frequency that is twice as geat as the original Nyquist frequency. Also, the PSD of the new signal in the upper half of the frequency 146 1.0 " (a) z < (9 0.0 °'° FREQUENCY °"" (b) T E U) 2 E E F “1 FLIGHT TIME —. Figure 7-8. - a) Frequency response of decimation filter. (b Filtered time-of- flight mass spectrum of xenon; filtered, then decn)mated by 10. 147 spectrum is the mirror image of the PSD in the lower half of the spectrum, as shown in Figure 7-9. Although this is not intuitively obvious, consider what would happen if every second point were removed. The result would be the original waveform, with its original PSD. If the upper half of the PSD of the interpolated signal were anything other than the mirror image of the lower half, this process of interpolation followed by decimation would result in a different PSD than the original waveform had. Obviously, placing alternating zeros in a data set is not what is desired in the process of interpolation by two. What we want is for the PSD of the signal that results from the interpolation process to be equal to the PSD of the signal that would have been obtained if the analog signal had been sampled at twice the sampling frequency. We presume that the original sampling frequency was adequate for representing the signal; that is, the PSD of the analog signal was nearly zero at frequencies higher than half the sampling frequency. If this is true, the upper half of the PSD of the interpolated signal should be nearly zero. Therefore, all we need to do is filter the signal that has the alternating zeros with a low-pass filter that has a cutoff frequency fc of 0.25 (half the Nyquist frequency) to obtain an interpolated version of the original waveform. This process is shown in Figure 7-9. Figure 7-10 shows the results of interpolation by a factor of two applied to a calculated Gaussian profile with a FWHM of approximately 9 points. Figure 7-10a shows the zero-filled data set. Figure 7-10b shows what happens when a 21 point eighth order SG filter is used to filter the zero-filled data set. Figure 7-10c shows the results obtained with a WILP filter that has a cutoff frequency of 0.477. The frequency responses of the two filters are shown in Figure 7-11. The SG filter is clearly inadequate for this purpose. The ripples in the stopband of the 86 filter let through some high frequency components that show up as sinusoidal components in the interpolated version of the Gaussian profile. One can see from examining Figure 7-9 that inserting the alternating zeros into 148 ‘ ladd zeros amplitude jfflter I frequency Figure 7-9. Interpolation by a factor of two. 149 .. °.' (a) M‘s-aooooooooooooooe.o°w Figure 7-10. Interpolated Gaussian ofile: (a) before filtering, (b) results of Savitzky-Golay interpo ation filter, (c) results of windowed ideal low-pass interpolation filter. 150 1.0 (b) (8) GAIN \ A / FREQUENCY Figure 7-11. Frequency responses of decimation filters: a3. windowed ideal low-pass filter, f = 0.477, (b) 21 point e' order Savitzky- C Golay filter. 151 the data set causes the PSD to be as geat at the Nyquist frequency as it is at DC. In the previous chapter it was shown that the SG filters all have gains that are nonzero at the Nyquist frequency. Therefore any 86 filter would be expected to perform poorly as an interpolation filter. As an interpolation scheme, the filtering method described above has some advantages over other methods. The filtering method makes no assumptions about what the form of the data is; it only assumes that the PSD of the data is insignificant above the original Nyquist frequency. In this way it is similar to the Fourier "zoom transform" approach, where one pads the Fourier transform of a signal with zeros to double its length, and then inverse transforms to produce the interpolated data (7). This contrasts with polynomial interpolation, where it is assumed that the data can be described by a polynomial of low degee. Choice of the proper degee for the polynomial is often a difficult task. The convolution method is much faster to implement than polynomial interpolation, especially if digital signal microprocessors are used, because it only requires multiplication and addition. The convolution method is more versatile than the zoom transform. Through combinations of interpolation and decimation any rational amount of interpolation can be done, such as interpolation by 4 and decimation by 3 to increase the effective sampling rate by 4/3. The zoom transform is limited to interpolation and decimation by powers of two unless the much slower discrete Fourier transform is used in place of the fast Fourier transform. CONCLUSIONS In this chapter it has been shown that Savitzky-Golay filters are not adequate for all filtering purposes. The removal of correlated noise, decimation and interpolation can be done better by filters that are designed specifically for those tasks. The design of a 152 filter is not a complicated process; simple filters can be designed with the aid of calculator. While not everyone who uses filters should feel compelled to design them as well, it is hoped that this chapter has shown that a little extra effort can sometimes pay off. 95"?!” >3 153 CHAPTER 7 REFERENCES Hamming, R.W. Digital Filters, 2nd ed., Prentice-Hall: Englewood Cliffs, NJ, 1983. Williams, C.S. Designing Digital Filters, Prentice-Hall: Englewood Cliffs, NJ, 1986. Ramirez, R.W. The FFT, Prentice-Hall: Englewood Cliffs, NJ, 1985. Marchand, P.; Marmet, L. Rev. Sci. Instrum I977, 48, 512. Stults, C.L.M.; Wade, A.P.; Crouch, S.R. Anal. Chim Acta 1987, 192, 301. Crochiere, R.E.; Rabiner, LR. Multimte Digital Signal Processing, Prentice-Hall: Englewood Cliffs, NJ, 1983. Francl, T.J.; Hunter, R.L.; McIver, R.T., Jr.;Anal. Chem 1983, 55, 2094. Patton, CJ. Ph. D. Dissertation, Michigan State University, 1982. Erickson E.; private communiation.