w. - o n- . aw - «I Ag; a ‘ €33; J 5%}. m G?!" i,L ' 1 1133;141:1514“; . E" Jr: . 4L. . 1.: ”,2“. _.._ ffiz fit“ 111i“ 111 'V 1“; 1.11.1115.“ 11731111 ”*1?“ 11:11?“ 2%}? 111%,? $1. $111111 "11%;: M 11%; a fit 1‘ .r “Lu 11111111! 5} 1 V .5rLs “I - ("11111131 1.111% .fiflif . ‘i‘mm .; x a r" '. 133394..” V- ‘1 ._' M- . ' 'I “ .‘ . , -. | '.1 11.0 . I'u ‘.4' .v‘ . . as . . ~ ‘- ,. . . :3:-.r' ’ ,. a”; .. .. . _. -L . . . ..... - . . . ‘. ' ‘1 “dc 5M... 4 . . 1:. -_ w«.|-vr‘r - 5'. ”3:79: 4 .._., . V"-..-Tfl .. u.-. - “n.3,: . .r:."~. ' A. h -. :3! .2' wk .4 -w‘ . ‘5'. ' . "17. L v- ";v- . ‘14- ;z': . .‘F-TTK'. '. u'flx . w .. . '23? .... . ._....- 5'73? -.,A .Vrgyflig: . . . . W- ‘2." a , Ff}, ‘ J. m i I: my 3?. ;w wm «- no 1‘) “'.;' “‘3'." «an... «- man r r » a... 1~§ - r u p ‘: . ‘, ._a..‘ ”Ev"- ~."~ E“. “ii $51k; ': y.- .n-chobv 1' 35‘"? I...“ I” - n 7:. W 3'32, :2. V w .fi. -_ ..—~.- ‘vy . o}..:a-: 1m .0- - w-..— Wat’s» 3‘35!" Cr;— '1" , M 1.1 n v o..- . P,“ .-- h... . an: o‘u— :' .. ‘35P?“ "M ' {fin are? an ,. .. ‘I‘A ' .-\ n . mm??? Mr. ‘ “Ha. 4: r0.- 3 .. ,, 2‘ 51¢ vrv - urn «w .-‘...,‘-~ urn-n- - 3 4: q .1 _ '1'.“ “21.... “a... ..‘,.—< ~ - . o :lifli .. 1‘ _ 1 . ‘3;‘¥¥3i-"._ 5 7‘?) 1 "A". E 1-! .1 ' ? 1111131111115 :1 * l ‘ gy $13}I.l‘ —-.o<. 1 1 ' ”111‘ ‘ 1. .1 In , . ‘ grim-12w: NM 51;»ng , 3;: g . d" M43 3-4'” . 1 1 ‘1' ‘:1 11' "‘11:”: “‘1‘“! 11:11:11 .1233 1211' 1,1; :hfifiég‘ 13’1931 :_ “ 1 . -. ,' ., 114:1” 112.1”1313h2'iil1ffldv 1 if} DEE "1:339" ‘ . z. . 11? i" 1t;‘1:}--1.-:: 1.111.1‘1'11'11111'1511 ' 311$ 1111111 I" .-.‘ -‘ mf‘“ ~23 1" 1'::11:11li§1i'1';l’3 a 11111“i1-';{5139‘ . “ 51;?“ , “11 E g‘ ' my. H‘vfirwixgn ..'§.‘.‘.C)53, a recursive algorithm for extrapolating slow kinetic profiles and instrumental responses“, and an error-compensating algorithm capable of determining the reaction order, rate constants, and concentrations of an system of general-order reactions55. This latter algorithm, however, requires modification for use with a first order system. 1.2.1.3. Kinetic Determinations using Nonlinear Regression In cases where the reaction order and rate constants of a reaction system are well known, nonlinear regression methods can be used to determine the concentration of one or more analytes. These regression methods have been applied to a variety of chemical systemsS‘r“. Some workers have also described the use of multidimensional nonlinear least squares fits of kinetic- 18 spectrophotometric data in cases where both the rate constants and absorptivities of the analytes are known61v65v66. 1.2.1.4. Kinetic Determinations using the Kalman Filter The Kalman filter67 is a recursive algorithm well suited to use with kinetic data“. Its use for such data has been reviewed”. It has also been found to be useful for enzymatic kinetic determinations"). Use of the Kalman filter requires knowledge (or at least a good estimate) of the rate constants. It presumes a reaction order. The application of the extended Kalman filter to kinetic-spectrophotometric data has been described by Quencer and Crouch“. It has been used in binary systems with first order kinetics72 as well as systems of consecutive reactions73. Parallel Kalman filter networks have been applied to kinetic determinations by Wentzell74. The Kalman filter and extended Kalman filter have been compared to other data processing techniques in a number of reviews37»33»4‘. 1.2.2. Kinetic Determinations using Soft Modeling Techniques Soft modeling techniques have the obvious advantage that they do not presume a model that may not fit the system being studied. They do, however, have the disadvantage that models they generate are often simply empirical and so have no physical meaning beyond their predictive ability. Several of these soft modeling techniques are described below. 19 1.2.2.1. Kinetic Determinations using Factor Analysis While not strictly a soft modeling technique (first order kinetics are usually assumed), factor analysis is closely related to some other multivariate calibration techniques, and so in included in this section. In systems where there is at least a moderate degree of spectral and kinetic resolution between the analytes, factor analysis has been used to determine the rate constants, absorbance spectra, and concentrations of mixtures of analytes75'31. 1.2.2.2. Kinetic Determinations using MLR, PCR and PLS Multivariate calibration techniques are finding broad application for kinetic determinations. This use has been reviewed in several papers36o33’39’41’42. Several of these applications are summarized here. Gallium and aluminum react with 4-(2-pyridylaxo) resorcinol (PAR) to produce products with very similar spectra. The ratio of the rate constants is kN/kca = 3.67. Using a stopped-flow, flow injection (FI) system with diode array detection, Blanco, et a1.82 determined mixtures of Ga and A1 with an error of less than 10%. In other work, O-O’-bis-(2-aminoethyl) ethylene glycol-N,N,N’,N’ tetraacetic acid (EGTA) complexes of Fe(II), Co(II), and Zn(II) were reacted with PAR"). These metal ions react with similar kinetics to form products with very similar kinetic profiles. This experiment was performed in a stopped-flow FI system with diode array detection. PCR and PLSR were used to determine Fe, Co and Zn successfully. The kinetics can be complicated by performing the 20 experiment in two steps in a flow system. If Co, Fe, and Zn are directly injected into the flow system, where they first react with EGTA and then with PAR, the kinetics of the Co and Zn are essentially the same as for the case where the EGTA complexes are directly injected. Iron(II), however, reacts slowly with EGTA and thus the kinetics associated with the formation of the Fe-PAR complex are significantly altered (in a non-linear fashion). Both methods (PCR and PLSR) were used to determine Co, Fe, and Zn using data collected in this second manner. Almost identical results were obtained with PCR and PLSR. They predicted the concentration of Zn and Co with good accuracy, but performed less well in determining Fe. In other work, Havel and coworkers determined vanadium and cobalt by PLSR using kinetic data”. The reaction studied was that of V and Co with the TrAMeR reagent (4-(1’H—l’,2’,4’-triazolyl-3’-azo)-2-methylresorcinol). The reaction was monitored at 60 3 intervals for 30 minutes at five wavelengths between 500 and 540 nm. The average relative percent error was 4%. In the same paper, a stopped Fl determination of Zn, Co, and Fe was described. The average error associated with this determination was also about 4%. Lopez-Cueto and coworkers84 have described the determination of aminophenol isomers. These authors used PLSR with kinetic-spectrophotometric data that were acquired with a diode array detector. The reaction studied was one that required that the reagent not be present in excess. Also, the concentration of 21 each isomer influenced the reaction rate of the others. In spite of the inherent kinetic non-linearity, acceptable results were obtained. Havel and coworkers85 reported on the kinetic-spectrophotometric determination of europium, terbium and lanthanum using PLSR. Binary mixtures of the metal ions reacted with Xylenol Orange to produce similar spectra. Acceptable errors were obtained (0.2-4%). The authors noted that the PLSR algorithm required at least four latent variables for a satisfactory fit. They also reported that, while excellent results were obtained with binary mixtures, ternary mixtures could not be resolved with acceptable error levels. A variety of other kinetic determinations have been achieved using multivariate calibration techniqueslo’ll‘“'32-‘06. 1.2.2.3. Kinetic Determinations using Artificial Neural Networks Artificial neural networks (ANNs) have been used for a variety of kinetic determinationslo’l3959940910741°. In general, it has been found that ANNs are similar to PLSR and PCR in their predictive ability in most cases. Artificial neural networks require more rigorous and lengthy calculations. They also require larger calibration sets. In most cases, the advantage (if any) of using an ANN is more than outweighed by these drawbacks. ANNs do, however, achieve superior results in cases where the data is nonlinear or in other ways ill-behaved10-13’95. 1.2.2.4. Kinetic Determinations using Multiway techniques Multiway techniques have been applied to kinetic determinations in only a handful of papers. The use of GRAM for kinetic-spectrophotometric data111 has 22 been investigated. Several papers have described initial studies of the application of nPLS to kinetic-spectrophotometric datam’m. A number of papers have reported the use of multivariate curve resolution for kinetic-spectrophotometric determinations1 14" 16. 1.3. (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) REFERENCES Wise, B. M.; Gallagher, N. B. PLS_Toolbox 2.0 User Manual ; Eigenvector Research, Inc.: Manson, WA, 1998. Kramer, R. Chemometric Techniques for Quantitative Analysis ; Marcel Dekker, Inc.: New York, 1998. Beebe, K. R.; Pell, R. J .; Seasholtz, M. B. Chemometrics: A practical guide ; John Wiley and Sons: New York, 1998. Booksh, K. S.; Kowalski, B. R. "Theory of analytical-chemistry" Anal. Chem. 1994, 66, A782—A79l. Sanchez, E. S.; Kowalski, B. R. "Tensorial calibration: I. First-order .. calibration" J. Chemometr. 1988, 2, 247-264. Martens, H.; Naes, T. Multivariate Calibration ; Wiley: New York, 1989. Malinowski, E. R.; Howery, D. G. Factor Analysis in Chemistry ; John Wiley and Sons: New York, 1980. Jolliffe, I. T. Principal Component AnalysisSpringer Series in Statistics, ; / Springer-Verlag: New York, 1986. Brereton, R. G. "Deconvolution of mixtures by factor-analysis" Analyst 1995, 120, 2313-2336. Blanco, M.; Coello, J.; Iturriaga, H.; Maspoch, S.; Redon, M. "Artificial neural networks for multicomponent kinetic determinations" Anal. Chem. 1995, 67, 4477-4483. / / / Geladi, P.; Kowalski, B. R. "Partial least-squares regression - a tutorial" / Anal. Chim. Acta 1986, 185, 1-17. 23 (12) (13) (14) (15) (16) (17) (18) (19) (20) (21) (22) (23) Geladi, P.; Kowalski, B. R. "An example of 2-block predictive partial least- squares regression with simulated data" Anal. Chim. Acta 1986, 185, 19-32. Blanco, M.; Coello, J .; Iturriaga, H.; Maspoch, S.; Redon, M.; Villegas, N. "Artificial neural networks and partial least squares regression for pseudo- first-order with respect to the reagent multicomponent kinetic- spectrophotometric determinations" Analyst 1996, 121, 395-400. Wise, B. M.; Ricker, N. L. "Identification of finite impulse-response models with continuum regression" J. Chemometr. 1993, 7, 1-14. Stone, M.; Brooks, R. J. "Continuum regression - cross-validated 4 sequentially constructed prediction embracing ordinary least-squares, partial least- squares and principal components regression" J. R. Stat. Soc. Ser. B-Methodol. 1990, 52, 237-269. Kalivas, J. H. "Interrelationships of multivariate regression methods using eigenvector basis sets" J. Chemometr. 1999, 13, 111-132. Malinowski, E. R. "Automatic window factor analysis - a more efficient method for determining concentration profiles from evolutionary spectra" J. Chemometr. 1996, 10, 273-279. Harvey, D. T.; Bowman, A. "Factor-analysis of multicomponent samples" J. Chem. Educ. 1990, 67, 470-472. Zupan, J.; Gasteiger, J. "Neural networks: A new method for solving chemical problems or just a passing phase?" Anal. Chim. Acta 1991, 248, l- 30. Despagne, F.; Massart, D. L. "Neural networks in multivariate calibration" Analyst 1998, 123, 157R-l78R. Sanchez, E.; Kowalski, B. R. "Generalized rank annihilation factor- analysis" Anal. Chem. 1986, 58, 496-499. Sanchez, E.; Ramos, L. S.; Kowalski, B. R. "Generalized rank annihilation method .1. Application to liquid- chromatography diode-array ultraviolet detection data" Journal of Chromatography 1987, 385, 151-164. Sanchez, E. S.; Kowalski, B. R. "Tensorial calibration: II. Second-order calibration" J. Chemometr. 1988, 2, 265-280. 24 L/ (24) (25) (26) (27) (28) (29) (30) (31) (32) (33) (34) (35) (36) (37) Ramos, L. S.; Sanchez, E.; Kowalski, B. R. "Generalized rank annihilation ’ method .2. Analysis of bimodal chromatographic data" Journal of Chromatography 1987, 385, 165-180. Bro, R. "PARAFAC. Tutorial and applications" Chemometrics Intell. Lab. Syst. 1997, 38, 149-171. Bro, R. "The N-way on-line course on PARAFAC and PLS" http://newton/foodsci.kvl.dk/n-way 1998. Harshman, R. A. "Foundations of the PARAFAC procedure: Model and conditions for an ’explanatory’ multi-mode factor analysis" UCLA Working Papers in Phonetics 1970, I6, 1. Bro, R. "Multiway calibration. Multilinear PLS" J. Chemometr. 1996, 10, 47-61. Perez-Bendito, D.; Silva, M. Kinetic Methods in Analytical Chemistry ; Ellis Horwood: Chichester, 1988. Perez-Bendito, D. "Inorganic differential kinetic-analysis - a review" Analyst 1984, 109, 891-899. Perez-Bendito, D. "Approaches to differential reaction-rate methods" Analyst 1990, 115, 689-698. Mottola, H. A. Some Kinetic Aspects of Analytical Chemistry ; Wiley: New York, 1988. Crouch, S. R. "Kinetic methods for intelligent automation" Chemometrics Intell. Lab. Syst. 1990, 8, 259-273. Crouch, S. R. "Trends in kinetic methods of analysis" Anal. Chim. Acta 1993, 283, 453-470. Crouch, S. R. "Kinetic methods of analysis - how do they rate" J. Chin. Chem. Soc. 1994, 41, 221-229. Crouch, S. R.; Cullen, T. F.; Scheeline, A.; Kirkor, E. S. "Kinetic determinations and some kinetic aspects of analytical chemistry" Anal. Chem 1998, 70, 53R-106R. Quencer, B. M.; Crouch, S. R. "Multicomponent kinetic methods" Crit. Rev. Anal. Chem. 1993, 24, 243-262. 25 (38) (39) (40) (41) (42) (43) (44) (45) (46) (47) (48) Perez-Bendito, D.; Silva, M. "Recent advances in kinetometrics" Trac- Trends Anal. Chem. 1996, 15, 232-240. Otto, M. "Chemometrics in kinetic-analysis" Analyst 1990, 115, 685-688. Cerda, V.; Cladera, A.; Estela, J. M. "Multicomponent techniques as applied to kinetic analysis" Quim. Anal. 1996, 15, 341-350. Xie, Y. L.; BaezaBaeza, J. J.; RamisRamos, G. "A comparative study of several chemometric methods applied to the treatment of two-way kinetic- spectral data for mixture resolution" Anal. Chim. Acta 1996, 321 , 75-95. Cullen, T. F.; Crouch, S. R. "Multicomponent kinetic determinations using multivariate calibration techniques" Mikrochim. Acta 1997, 126, 1-9. Love, M. D.; Pardue, H. L. "Systematic comparison of data-processing options for kinetic- based, single-component determinations of noncatalysts: Effects of random noise" Anal. Chim. Acta 1996, 326, 95-106. Love, M. D.; Pardue, H. L. "Systematic comparison of data-processing options for kinetic- based single-component determinations of non-catalysts .1. Review, systematic classification, mathematical descriptions, performance-characteristics and perspectives" Anal. Chim. Acta 1994, 299, 195-208. Love, M. D.; Pardue, H. L. "Systematic comparison of data-processing options for kinetic- based single-component determinations of non-catalysts .2. One- rate, 2-point/fixed-time, 2-rate, 3-point/fixed-time options" Anal. Chim. Acta 1994, 299, 209-218. Wentzell, P. D.; Crouch, S. R. "Comparison of reaction-rate methods of analysis for systems following lst—order kinetics" Anal. Chem. 1986, 58, 2855-2858. Boschreig, F.; Campinsfalco, P.; Sevillanocabeza, A.; Herraezhemandez, R.; Molinslegua, C. "Development of the h-point standard-additions method for ultraviolet visible spectroscopic kinetic-analysis of 2- component systems" Anal. Chem. 1991, 63, 2424—2429. Marquez, M.; Silva, M.; Perez-Bendito, D. "Continuous addition of reagent technique - a new approach to differential reaction-rate methods" Anal. Chim. Acta 1990, 239, 221-227. 26 (49) (50) (51) (52) (53) (54) (55) (56) (57) (58) (59) Velasco, A.; Silva, M.; Perez-Bendito, D. "Processing analytical data obtained from 2nd-order reactions by using continuous reagent addition" Anal. Chem. 1992, 64, 2359-2365. Carreto, M. L.; Lunar, L.; Rubio, S.; Perez-Bendito, D. "Simultaneous spectrophotometric determination of o-cresol and m-cresol in urine by use of the kinetic wavelength-pair method" Analyst 1996, 121, 1647-1652. Carreto, M. L.; Lunar, L.; Rubio, S.; Perez-Bendito, D. "Simultaneous spectrophotometric determination of chlorpromazine, perphenazine and acetopromazine by use of the kinetic wavelength pair-method" Anal. Chim. Acta 1997, 349, 33-42. Schechter, I. "Error-compensated kinetic determinations by detecting the intermediate product in successive reactions, without prior knowledge of reaction constants" Anal. Chem. 1991, 63, 1303-1307. Schechter, I.; Schroder, H. "Error-compensated kinetic determinations in systems of mixed 1st-order and 2nd-order reactions, without prior knowledge of reaction constants" Anal. Chem. 1992, 64, 325-329. Schechter, I. "Predictive mode of kinetic-analysis and transient detection responses - evaluation of a recursive algorithm" Anal. Chem. 1992, 64, 2610-2614. Schechter, I. "Simultaneous determination of mixtures by kinetic-analysis of general-order reactions" Anal. Chem. 1992, 64, 729-737. Cost, J. R. "Non-linear regression least-squares method for determining relaxation-time spectra for processes with lst-order kinetics" J. Appl. Phys. 1983, 54, 2137-2146. Gomez, E.; Cladera, A.; Estela, J. M.; Cerda, V. "Multidata treatment applied to the simultaneous resolution of catechol resorcinol mixtures by kinetic enzymatic processes" Talanta 1993, 40, 1601-1607. Diaz, A. N.; Garcia, J. A. G. "Nonlinear multicomponent kinetic-analysis for the simultaneous stopped-flow determination of chemiluminescence enhancers" Anal. Chem. 1994, 66, 988-993. Cai, Q. Y.; Wang, R. G.; Wu, L. Y.; Nie, L. H.; Yao, S. Z. "A nonlinear regression model applied to kinetic studies of ester hydrolysis with a surface acoustic wave sensor" Talanta 1996, 43, 699-705. 27 (60) (61) (62) (63) (64) (65) (66) (67) (68) (69) (70) Gill, A.; Leatherbarrow, R. J.; Hoare, M.; PollardKnight, D. V.; Lowe, P. A.; Fortune, D. H. "Analysis of kinetic data of antibody-antigen interaction from an optical biosensor by exponential curve fitting" J. Biotechnol. 1996, 48, 117-127. Havel, J.; Cuesta, F.; Gonzalez, J. L.; Canedo, M. D. "Multicomponent kinetic analysis and kinetic studies .1. Multipurpose KILET(94) non-linear least squares program" J. Anal. Chem. 1996, 51, 100-105. Kuroda, Y.; Kawashima, A.; Ogoshi, H. "Computer analyses of complex kinetics containing equilibrium processes. Example of application for unusual atropisomerization of a tetraphenylporphyrin derivative" Chem. Lett. 1996, 57-58. Roecker, L.; Lee, S. M.; Liu, L. W.; Sun, L. J.; Zaerpoor, K. "Analysis of first-order kinetic data by a differential technique" Int. J. Chem. Kinet. 1997, 29, 333-338. Shultz, L. L.; Nieman, T. A. "Stopped-flow analysis of Ru(bpy)(3)(3+) cherrriluminescent reactions" J. Biolumin. Chemilumin. 1998, 13, 85—90. Chau, F. T.; Mok, K. W. "Multiwavelength analysis for a 1st-order consecutive reaction" Comput. Chem. 1992, 16, 239-242. Endo, M.; Sasaki, 1.; Abe, S. "Kinetic spectrophotometric determination of iron oxidation- states in geological-materials" F resenius J. Anal. Chem. 1992, 343, 366-369. Brown, S. D. "The Kalman filter in analytical-chemistry" Anal. Chim. Acta 1986, 181, 1-29. Wentzell, P. D.; Karayannis, M. 1.; Crouch, S. R. "Simultaneous kinetic determinations with the Kahnan filter" Anal. Chim. Acta 1989, 224, 263- 274. Xiong, R.; Velasco, A.; Silva, M.; Perez-Bendito, D. "Performance of the kalman filter algorithm in differential reaction-rate methods" Anal. Chim. Acta 1991, 251, 313-319. Lin, D. Z.; Tan, H. W.; Bao, L. L.; Nie, L. H.; Yao, S. Z. "Study on a modified algorithm - iterative Kalman filter and fits application in enzymatic kinetics" Analyst 1995, 120, 2573-2578. 28 (71) (72) (73) (74) (75) (76) (77) (73) (79) (80) (81) Quencer, B. M.; Crouch, S. R. "Extended Kalman filter for multiwavelength, multicomponent kinetic determinations" Analyst 1993, 118, 695-701. Quencer, B. M.; Crouch, S. R. "Multicomponent kinetic determination of lanthanides with stopped-flow, diode-array spectrophotometry and the extended Kalman filter" Anal. Chem. 1994, 66, 458-463. Mok, K. W.; Chau, F. T. "Application of the extended Kalman filter for analysis of a consecutive first-order reaction" Trac-Trends Anal. Chem. 1996, 15, 170-174. Wentzell, P. D.; Vanslyke, S. J. "Parallel Kalman filter networks for kinetic methods of analysis" Anal. Chim. Acta 1992, 257, 173-181. Cladera, A.; Gomez, E.; Estela, J. M.; Cerda, V. "Resolution of simultaneous kinetic spectrophotometric processes by factor-analysis" Anal. Chem. 1993, 65, 707-715. Tam, K. Y.; Chau, F. T. "Simultaneous multiwavelength study of the reduction reaction of chromic-acid with oxalic-acid using a photodiode- array spectrophotometer" Spectr. Lett. 1993, 26, 1195-1212. Tam, K. Y.; Chau, F. T. "Multivariate study of kinetic data for a 2-step consecutive reaction using target factor-analysis" Chemometrics Intell. Lab. Syst. 1994, 25, 25-42. Liang, X. H.; Andrews, J. E.; deHaseth, J. A. "Resolution of mixture components by target transformation factor analysis and determinant analysis for the selection of targets" Anal. Chem. 1996, 68, 378-385. Garland, M.; Visser, E.; Terwiesch, P.; Rippin, D. W. T. "On the number of observable species, observable reactions and observable fluxes in chemometric studies and the role of multichannel integration" Anal. Chim. Acta 1997, 351, 337-358. Fister, J. C.; Harris, J. M. "Resolving component overlap in multiwavelength kinetic spectroscopy: Application to Raman scattering from intermediates in triplet-state photoreactions" Photochem. Photobiol. 1997, 65, 47-56. Amrhein, M.; Srinivasan, B.; Bonvin, D. "Target factor analysis of reaction data: Use of data pre- treatment and reaction-invariant relationships" Chem. Eng. Sci. 1999, 54, 579-591. 29 (82) (83) (84) (85) (86) (87) (88) (89) (90) Blanco, M.; Coello, J.; Iturriaga, H.; Maspoch, S.; Riba, J.; Rovira, E. "Kinetic spectrophotometric determination of Ga(III)-Al(III) mixtures by stopped-flow injection-analysis using principal component regression" Talanta 1993, 40, 261-267. Havel, J.; Jimenez, F.; Bautista, R. D.; Leon, 1. J. A. "Evaluation of multicomponent kinetic-analysis data by a partial least-squares calibration method" Analyst 1993, 118, 1355-1360. Lopez-Cueto, G.; Maspoch, S.; Rodriguez-Medina, J. F.; Ubide, C. "Simultaneous kinetic spectrophotometric determination of o-, m- and p- aminophenol using partial least squares calibration" Analyst 1996, 121, 407-412. Fraga, J. M. G.; Abizanda, A. I. J.; Moreno, F. J.; Leon, J. J. A.; Havel, 1. "Kinetic spectrophotometric resolution of mixtures of europium, terbium, and lanthanum in a micellar medium using multivariate calibration" Microchem J. 1996, 54, 32-40. Havel, J .; Cuesta, F.; Jancar, L. "PLS self-calibration method for the evaluation of rate constants and or initial concentrations in multicomponent kinetic-analysis" React. Kinet. Catal. Lett. 1993, 49, 189-195. Blanco, M.; Coello, J.; Iturriaga, H.; Maspoch, S.; Riba, J. "Kinetic spectrophotometric method for analyzing mixtures of metal-ions by sopped-flow injection-analysis using partial least-squares regression" Anal. Chem. 1994, 66, 2905-2911. Baxter, P. J.; Christian, G. D.; Ruzicka, J. "Multivariate approach for the simultaneous determination of total biomass and glucose from a yeast fermentation by sequential injection-analysis" Chem. Anal. 1995, 40, 455- 471. Blanco, M.; Coello, J .; Iturriaga, H.; Maspoch, S.; Redon, M. "Partial least- squares regression for multicomponent kinetic determinations in linear and nonlinear-systems" Anal. Chim. Acta 1995, 303, 309-320. Garcia, J. M.; Jimenez, A. 1.; Arias, J. J.; Khalaf, K. D.; Moralesrubio, A.; Delaguardia, M. "Application of the partial least-squares calibration method to the simultaneous kinetic determination of propoxur, carbaryl, ethiofencarb and formetanate" Analyst 1995, 120, 313-317. 30 / /-. (91) (92) (93) (94) (95) (96) (97) (98) (99) On, Z. G; Wang, X. D. "The application of principal component regression on simultaneous multicomponent determinations through a single catalytic kinetic run" Talanta 1995, 42, 205-210. Lupu, M.; Todor, D. "A singular-value decomposition based algorithm for multicomponent exponential fitting of NMR relaxation signals" Chemometrics Intell. Lab. Syst. 1995, 29, 11-17. Meras, I. D.; Mansilla, A. E.; Lopez, F. S. "Simultaneous kinetic spectrophotometric determination of 2- furfuraldehyde and 5- hydroxymethyl-Z-furfuraldehyde by application of a modified Winkler’s method and partial least- aquares calibration" Analyst 1995, 120, 2567- 2571. Wang, X. D.; Gu, Z. C. "Generalized-method for simultaneous multicomponent determinations through a single catalytic kinetic run by using the rate spectrum" Analyst 1995, 120, 1839-1842. Blanco, M.; Coello, J .; Iturriaga, H.; Maspoch, S.; Redon, M.; Rodriguez, J. F. "Artificial neutral networks and partial least-squares regression for second-order multicomponent kinetic determinations" Quim. Anal. 1996, 15, 266-275. de la Guardia, M.; Khalaf, K. D.; Hasan, B. A.; Morales-Rubio, A.; Arias, J. J .; Garcia-Fraga, J. M.; Jimenez, A. I.; Jemenez, F. "Simultaneous kinetic spectrophotometric determination of five phenolic compounds by reaction with p-aminophenol, using partial least squares data treatment" Analyst 1996, 121, 1321-1326. Khalaf, K. D.; MoralesRubio, A.; DelaGuardia, M.; Garcia, J. M.; Jimenez, F .; Arias, J. J. "Simultaneous kinetic determination of carbamate pesticides after derivatization with p-aminophenol by using partial least squares" Microchem J. 1996, 53, 461-471. Lopez-Cueto, G.; Rodriguez-Medina, J. F.; Ubide, C. "Individual kinetic determinations using partial least squares calibration" Analyst 1997, 122, 519-523. Xing, W. L.; He, X. W. "Kinetic determination of organic vapor mixtures with single piezoelectric quartz crystal sensor using artificial neural networks and partial least squares" Chem. Lett. 1996, 1065-1066. 31 (100) (101) (102) (103) (104) (105) (106) (107) (108) (109) Xing, W. L.; He, X. W. "Kinetic determination of organic vapor mixtures with single piezoelectric quartz crystal sensor using artificial neural networks" Talanta 1997, 44, 959-965. Blasco, F.; Medina-Hemandez, M. J .; Sagrado, S. "Use of pH gradients in continuous-flow systems and multivariate regression techniques applied to the determination of methionine and cysteine in pharmaceuticals" Anal. Chim. Acta 1997, 348, 151-159. Blasco, F.; Medina-Hernandez, M. J .; Sagrado, S.; Fernandez, F. M. "Simultaneous spectrophotometric determination of calcium and magnesium in mineral waters by means of multivariate partial least-squares regression" Analyst 1997, 122, 639-643. Sans, D.; Nomen, R.; Sempere, J. "Interactive self-modelling of chemical reaction systems using multivariate data analysis" Comput. Chem. Eng. 1997, 21, $631-$636. Izquierdo, A.; Lopez-Cueto, G.; Medina, J. F. R.; Ubide, C. "Simultaneous determination of niobium and tantalum with 4-(2- pyridylazo) resorcinol using partial least squares regression and artificial neural networks" Quim. Anal. 1998, 17, 67-74. Kappes, T.; Lopez-Cueto, G.; Rodriguez-Medina, J. F.; Ubide, C. "Improved selectivity in multicomponent determinations through interference modelling by applying partial least squares regression to kinetic profiles" Analyst 1998, 123, 2071-2077. Peralta-Zamora, P.; Kunz, A.; Nagata, N.; Poppi, R. J. "Spectrophotometric determination of organic dye mixtures by using multivariate calibration" Talanta 1998, 47, 77-84. Ventura, S.; Silva, M.; Perez-Bendito, D.; Hervas, C. "Multicomponent kinetic determinations using artificial neural networks" Anal. Chem. 1995, 67, 4458-4461. Galvan, I. M.; Zaldivar, J. M.; Hernandez, H.; Molga, E. "The use of neural networks for fitting complex kinetic data" Comput. Chem. Eng. 1996, 20, 1451-1465. Ventura, 8.; Silva, M.; Perez-Bendito, D.; Hervas, C. "Estimation of parameters of kinetic compartrnental models by use of computational neural networks" J. Chem. Inf Comput. Sci. 1997, 37, 517-521. 32 i.‘\ (110) (111) (112) (113) (114) (115) (116) Ventura, 8.; Silva, M.; Perez-Bendito, D.; Hervas, C. "Computational neural networks in conjunction with principal component analysis for resolving highly nonlinear kinetics" J. Chem. Inf. Comput. Sci. 1997, 37, 287-291. Xie, Y. L.; BaezaBaeza, J. J.; RamisRamos, G. "Second-order tensorial calibration for kinetic spectrophotometric deterrrrination" Chemometrics Intell. Lab. Syst. 1996, 32, 215-232. Pettersson, A. K.; Karlberg, B. "Simultaneous determination of orthophosphate and arsenate based on multi-way spectroscopic-kinetic data evaluation" Anal. Chim. Acta 1997, 354, 241-248. Xie, Y. L.; Baezabaeza, J. J .; RamisRamos, G. "Kinetic spectrophotometric resolution of binary-mixtures using 3-way partial least-squares" Chemometrics Intell. Lab. Syst. 1995, 27, 211-220. Tauler, R.; Smilde, A. K.; Henshaw, J. M.; Burgess, L. W.; Kowalski, B. R. "Multicomponent determination of chlorinated hydrocarbons using a reaction-based chemical sensor .2. Chemical speciation using multivariate curve resolution" Anal. Chem. 1994, 66, 3337-3344. Saurina, J .; Hemandez-Cassou, S.; Tauler, R.; Izquierdo-Ridorsa, A. "Multivariate resolution of rank-deficient spectrophotometric data from first-order kinetic decomposition reactions" J. Chemometr. 1998, 12, 183- 203. Saurina, J.; Hemandez-Cassou, S.; Tauler, R. "Multivariate curve resolution and trilinear decomposition methods in the analysis of stopped-flow kinetic data for binary amino acid mixtures" Anal. Chem. 1997, 69, 2329-2336. 33 CHAPTER 2 INSTRUMENT DESIGN AND CHARACTERIZATION Never worry about theory as long as the machinery does what it’s supposed to do. --Robert A. Heinlein The data acquisition system used to collect the kinetic-spectrophotometric data discussed in this document consisted of a home-built stopped-flow apparatusl interfaced to a Tracor Northern (Model TN-6123) 512 element intensified diode array (Tracor Northern, Philadelphia, PA). Both the stopped-flow apparatus and the computerized interface between the stopped-flow and the diode array were modified for this research. These modifications and the subsequent characterization of the entire instrument are the subject of this chapter. 2.1. DESIGN OF DIODE ARRAY INTERFACE A new computerized interface between the diode array and the stopped- flow was designed and constructed. This interface controls the diode array and is responsible for sending timing signals to the array and acquiring the array’s output. The interface is composed of a peak track-and-hold circuit and a computer program that controls National Instruments LabPC+ and PC-TIO-lO data acquisition and timing boards. 34 A timing diagram describing the signals sent to and from the array is shown in Figure 2-1. Trigger Scan START BEOS STB I J Pulse Out A P Peak Track] | [— Read 1 Reset I L. 0 2 4 6 8 10 12118 Figure 2-1: Timing diagram of signals sent to and from the diode array. The Trigger signal is sent by an opto-interrupter circuit attached to the stopped- flow when a push has occurred and flow is stopped. When the computer sees Trigger go high, it sets Scan high. The length of time Scan stays high is determined by the length of the planned acquisition, i.e., on the number of scans required and on the scan frequency. START is gated off of Sean, and has a 35 frequency related to the desired scan rate. Upon receiving the START pulse, the diode array sets BEOS (Buffered End Of Scan) high. The STB (Scan Time Base) timing signal is gated off of BEOS and begins to send out pulses when BEOS goes high. The frequency of the STB pulses is related to the desired integration time for each diode. Each STB pulse results, after a short delay, in the appearance of a ~100ns pulse on the Pulse Out line. The pulses are the result of each diode’s accumulated charge being converted into a voltage pulse in turn. The magnitude of a pulse is related to the number of photons that impacted its diode during the integration time. The Peak Track line is the output of a peak track-and-hold circuit which is described in detail later in this section. This circuit tracks the pulse to its maximum and then holds at that level until reset. Six microseconds after an STB pulse, the Read line goes low, triggering the computer’s data acquisition board to initiate an acquisition of the Peak Track line. After a short delay, the Reset signal goes low and resets the peak track-and-hold circuit into tracking mode. Both the Read and the Reset signals are initialized by the START pulse. In reality, each scan of the array requires two START pulses and two STB pulse trains. The first “scan” is not read into memory, but is merely used to clear the array. The STB frequency is used to control the period between the clearing and reading of each diode. 36 .53.» .8 :23... 2: 5 8:33 93 239.53.: 24388:. «.933 95 8 98585.3 33:: >m+ can EoEEEa 3.8332? near—h 2: 3 93335.3 .332— >N~H .3595 2265.235 :3.— 95 he Engage ass—5:5 “TN 253 _ 33. >n+ 50 are .l m _l>n+ Gan—'10 < 9 an. 32. )n- I I -— _— — H §U_|>n+ .1 mom—.5 » see... :3.— all QZO 37 Figure 2-2 is a schematic of the peak track-and-hold circuit. The Pulse Out signal is inverted by an inverting amplifier. The second op amp has gain and inverts the signal again. The track-and-hold functionality is provided by a fast signal diode, the third op amp, and a high quality capacitor. The capacitor charges as the rising edge of a pulse comes down the Pulse Out line. As the pulse hits its maximum and begins to fall, the diode prevents the capacitor from discharging. The input impedence of the voltage follower immediately after the capacitor prevents discharge in that direction. The capacitor thus charges while a pulse is rising and the holds at the peak value after the pulse begins to fall. The capacitor is discharged by closing a switch that provides a low resistance path to ground. The switch is closed by a high on the Reset line. Reset is the output of a monostable multivibrator that clocks off of the STB pulse train. Shortly before the Reset goes high, a second monostable causes Read to go low, triggering the computer’s data acquisition board to initiate a sample-and-hold operation on the Peak Track signal. The LT1363 operational amplifiers used for the inverting amplifier, the inverting amplifier with gain, and the voltage follower are 70 MHz amplifiers with slew rates of 1000V/us in order to be able to respond to the very short duration pulses on the Pulse Out signal. The switching is done with a 4066 high speed quad bilateral switch. 38 2.2. REDESIGN OF STOPPED-FLOW OBSERVATION CELL AND OPTICS The optical path of the stopped-flow apparatus begins with a fiber optic bundle carrying light from a tungsten lamp to the observation cell. The light passes through the observation cell, where some is absorbed by the sample, and is collected by a second fiber optic bundle which carries it to the diode array detector. From mixer Fiber Optic Bundle to Diode Array Fiber Optic Bundle from Tungsten Lamp To stop syfinge Figure 2-3: Schematic of the observation cell. As seen in Figure 2-3, the fiber optic bundles are inserted into the block containing the observation cell and are butted against spherical sapphire ball lenses. These lenses are fitted against the openings at the ends of the observation cell and serve 39 both to seal the cell and to collimate and collect the light passed through the observation cell. They are held in place by custom-designed Delrin screws. In previous incarnations of the instrument, the fiber optic bundles did not extend into the block itself. Rather, a three centimeter quartz rod was inserted into the block and butted against a flat window. The rod was screwed into place with sufficient pressure to force the window to seal the cell. The fiber optics were touched to the quartz rod. Thus, the light had to pass from the fiber optic bundle, through the rod, through the window, and into the cell. This resulted in several interfaces where light could be lost due to reflection. The current scheme, in which the quartz rods have been eliminated and the fragile quartz windows replaced by extremely durable sapphire balls, is significantly simpler to operate and maintain. The light throughput is at least as good as the previous arrangement. 2.3. CHARACTERIZATION OF DATA ACQUISITION SYSTEM After the computerized interface was in place and the redesigned optical path had been implemented, a detailed characterization study of the instrument as a whole was performed. The delay time and maximum scan rate were determined. The mixing time was not determined for reasons discussed below. The signal-to- noise ratio of the instrument was measured, and the linear range was found. 2.3.1. Determination of Delay Time and Maximum Acquisition Rate Literature procedures2 were used in an attempt to determine the dead and mixing times of the instrument. An attempt was made to experimentally determine 40 the dead time of the stopped flow by observing the reaction of iron (III) with thiocyanate to form a colored productz. Due to instrumental and electronic limitations, the dead time was not determined; rather another parameter we shall call the delay time was calculated by extrapolating the reaction curve back to the initial absorbance of the reactant mixture. Delay time, then, is defined as the time for which the reactants are mixing in the observation cell but are unobserved, i.e., the time between the mixing of the sample and reagent and the collection of the first data point. If the first data point is defined to be acquired at time zero, the delay time is the negative of the time at which the extrapolated reaction curve reaches the absorbance of the reactants. For this system, the delay time was determined to be 75 ms. Previous work using this same stopped-flow apparatus1 had found the dead time of the stopped-flow itself be approximately 5 ms. The difference between the two values can be attributed to several factors. Those that contribute most to the difference are the delay between stoppage of flow and the triggering of the opto- interrupter circuit, the delay between the arrival of the Trigger signal and the generation of the first START pulse, and the need for one complete scan of all 512 diodes (to clear the array) before any data can be acquired. The mixing time of the stopped-flow apparatus was not accurately determined because it was shorter than the delay time of the instrument as a whole, and so not measurable using the experimental setup used for the other experiments. 41 The maximum rate at which spectra could be collected was determined, and found to be a function of the desired integration time. The integration time determines the frequency of the STB pulse train, and therefore the length of time necessary to complete one full scan of the array. Due to the maximum STB frequency (80 kHz) supported by the track-and-hold circuit and the lag between the end of the clearing scan of the array and the beginning of the data acquisition scan (determined in part by the speed of the PC running the interface program, in our case a 33 MHz 486SX), the minimum practical integration time is 13 ms. At this minimum integration time spectra can be acquired at 10.5 Hz. Many studies described in this work were performed with 35 ms integration times. At that integration time, 5.5 spectra can be acquired per second.. 2.3.2. Determination of signal/noise ratio The signal/noise ratio (S/N) for each wavelength was measured as the range of recorded intensity measurements for a series of spectra acquired from a static system. The average S/N was computed as the average of the S/Ns of all wavelengths between 500 and 700 nm (the range used in these experiments). These S/N measurements were carried out under a variety of circumstances. The S/N was found to be a function of percent of the radiation transmitted through the observation cell, intensifier gain, and integration time. The dark noise (the range of intensities measured when no light impinges on the array) is a major source of noise. Accordingly, attempts were made to minimize it by cooling the array. A Peltier thermoelectric cooler was used to bring 42 the array temperature to —5.9°C. The cooler decreased the dark signal by only 2%, but decreased the amplitude of the dark noise by 30%. For all measurements, the average dark signal at each wavelength was removed in a background subtraction step. 0.0 -O.5 - -1.0 - -1.5 - 109(0 IA) -2.0 - -2.5 - '3.0 T i I T r 1 O 0.5 1 1 .5 2 2.5 3 3.5 Absorbance Figure 2-4: Plot showing dependence of relative error in absorbance to absorbance. Figure 2-4 shows the relationship between absorbance (in this case the absorber was a neutral density filter that attenuates the light) and the relative error in absorbance. As is expected, log(oA/A) goes through a minimum at an absorbance of ~0.5. It can also be seen that log(oA/A) is fairly constant between absorbances of 0.2 and 1.0. For this reason all experiments were run under conditions where no absorbances greater than 1.0 were measured. 43 2.5 4 Iog(S/N) = (0.91 *log(is)) + 2.2 A 2 d z R2=O.999 \ SP, 15 - O) .9 1 _ 0.5 - 0 t l r I j t -2.5 -2 -1 .5 -1 -0.5 0 0.5 1 109(is) Figure 2-5: Dependence of SIN on absorbance. Figure 2-5 shows a linear relationship between the logarithm of the signal to noise ratio and the logarithm of the measured signal (is), where the signal level was varied by varying the light intensity The different measured intensities are due to the attenuation of the light with a series of neutral density filters. The measured slope was 0.91. For a shot noise limited system a slope of 0.5 is anticipated; when blank noise dominates a slope of unity is expected3. It can thus be inferred that the measurement is not shot noise limited, but over this region a mixture of shot noise and blank noise predominates. We can speculate that blank noise is dominant at the low end of the region and shot noise is dominant when light levels are at the high end of the region shown in Figure 2-53. 2.2 2.1 4 log(S/N)= (0.64*|og(is)) + 2.1 R2 = 0.994 log (S/N) .5 _L _l —L —L in b) K1 in to N 1 1 1 1 1 db on I o a: I 9 rs ('3 N o 0.2 1090s) Figure 2-6: Dependence of SIN on intensifier gain. Figure 2-6 shows a linear relationship between the logarithm of the signal to noise ratio and the logarithm of the measured signal (is) where the measured signal was varied by changing the intensifier gain at a constant integration time (13 ms, the minimum integration time) and constant light intensity. As gain increases, so does the measured signal amplitude. At this minimum integration time, the slope of the linear plot is 0.64. Blank noise contributes less and the system is more nearly shot noise limited when intensifier gain is increased. 45 2 1.95 - 1.9 ~ log(S/N)= (0.93*log(is)) +2.26 1.85 i R2=0.998 1.8 — 1.75 - 1.7 ~ 1.65 ~ 1.6 - 1.55 - 1.5 T i . . -O.8 -o.7 -0.6 -0.5 -o.4 -o.3 1090s) Figure 2-7: Dependence of SIN on integration time. Iog(S/N) Figure 2-7 shows the relationship between the logarithm of the signal to noise ratio and the logarithm of the measured signal (is) where the measured signal is a function of integration time at a constant intensifier gain (0.2, a very low gain) and constant light intenisty. As integration time increases, so does the signal. Here the measured slope is close (0.93) to unity. At low gains blank noise is important. 2.3.3. Determination of linear range The range over which absorbance measurements are linearly related to the absorption properties of samples in the optical path was determined in two ways. First, a series of neutral density filters with known absorbances across a broad region of the spectrum were inserted into the optical path. Absorbance as measured by the diode array at all wavelengths between 500 and 700 nm was compared to the known absorbance of the filters. 46 y = 0.98x - 0.0026 R2 = 0.997 Measured Absorbance N 0 I A i 1 0 1 2 3 4 Nominal Absorbance Figure 2-8: Linearity of measured absorbance with the known absorbance of a series of neutral density filters. The linear regression shown was calculated from all absorbances up to and including one absorbance unit. As shown in Figure 2-8, the array shows linearity for absorbances at least up to one absorbance unit. Above this point stray light causes a negative deviation from Beers law. As a second check of linearity, a series of chromium (III) solutions were passed through the observation cell of the stopped-flow and their spectra recorded. A calibration plot of absorbance at 569 nm versus concentration was generated and found to be linear to absorbances of greater than 1.50 (the absorbance of the solution with the highest concentration). 47 2.4. (1) (2) (3) REFERENCES Beckwith, P. M.; Crouch, S. R. "An automated stopped-flow spectrophotometer with digital sequencing for millisecond analyses" Anal. Chem. 1972, 44, 221-227. Stewart, J. E. "Flow deadtime in stopped-flow measurements" Dionex Application Notes , 1-4. Ingle, J. D., Jr.; Crouch, S. R. Spectrochemical Analysis ; Prentice Hall: Englewood Cliffs, NJ, 1988. 48 CHAPTER 3 INITIAL SIMULATION STUDIES: STUDY OF THE EFFECT OF EXPERIMENTAL VARIABLES The sciences do not try to explain, they hardly even try to interpret, they mainly make models. By a model is meant a mathematical construct which, with the addition of certain verbal instructions, describes observed phenomena. The justification of such a mathematical construct is solely and precisely that it is expected to work --John Von Neumann I do not fear computers. I fear the lack of them. --Isaac Asimov In order to understand the factors that limit the accuracy with which an analyte can be determined in a sample using kinetic-spectrophotometric data, a systematic study of the effect of an array of experimental parameters on the accuracy of a kinetic-spectrophotometric determination was performed. Because of the impracticality of performing the number of experiments needed to complete the study, simulated experiments were used. In these simulations experimental parameters were varied and the effect of these variations on the accuracy of a multicomponent (in this case, a two-component) kinetic-spectrophotometric determination was noted. 49 3.1. EXPERIMENTAL 3.1.1. Generation of Simulated Data Multicomponent kinetic-spectrophotometric data were simulated using a program written in MATLAB (The Math Works, Natick, Mass). The algorithm generates kinetic profiles for parallel second order reactions (first order in analyte and first order in reagent) by numerically solving the appropriate differential equations. The program mulgen_a which was used for this purpose is given in the appendix. For most of the simulation studies, synthetic spectra and rate constants were used to generate kinetic-spectrophotometric data. The absorbance data were generated by assuming that the analytes and reagent do not absorb in the spectral region of interest and that the absorption spectra of the reaction products can be modeled as overlapped Gaussian-shaped profiles. These synthetic spectra are shown later in this chapter. In all cases, adherence to Beer’s law was presumed for each component, and the total absorbance at each wavelength was assumed to be the sum of the absorbances of the components. 3.1.2. Data processing Time dependent spectra were collected in triplicate and averaged, i.e., each metal ion solution was reacted with PAR three times. The resulting three sets of kinetic- spectrophotometric data were averaged. Data were mean-centered (i.e., the mean of each variable vector was subtracted from each of its elements) before being 50 input to the appropriate algorithms. Multivariate calibration algorithms provided in the PLS_TOOLBOX (Eigenvector Technologies, Manson, WA.) and run in MATLAB were used to perform determinations. 3.2. METHODS FOR QUANTIFYING KINETIC AND SPECTRAL DIFFERENCES FOR TWO COMPONENTS In a kinetic-spectrophotometric determination, the accuracy with which an analyte can be determined in unknown samples depends on the degree of kinetic and spectral differentiation between the analyte and the other components of the sample. In order to examine the effect of an array of experimental parameters on both degree of kinetic and spectral differentiation and on the accuracy of a kinetic- spectrophotometric determination, it was necessary to develop means for quantifying the degree of kinetic and spectral differentiation between the analytes in a mixture. 3.2.1. Methods for quantifying kinetic differences In the following discussion, and indeed in most of this document, it is assumed that in kinetic or kinetic-spectrophotometric determinations analytes react with a common reagent to form different products at different rates. The plot of concentration of product vs. time (or of absorbance vs. time, if either the analyte or its product absorbs in the wavelength region of interest) is referred to as a kinetic profile. 51 The amount of kinetic differentiation between two components can be described by the ratio of the rate constants for the reactions of the analytes with the reagent. As the ratio of the rate constants increases, so does the difference between the rate constants and thus the difference in the kinetic profiles of the reactions of the analytes. The angle (6K) between the kinetic profiles of the two reaction products was used as a measure of the degree of kinetic differentiation between the analytes, and was calculated as the arccosine of the correlation between the kinetic concentration profiles of the reaction products. 6K = arccos [1%.] 0' K1 ' 0K2 where K and K2 are the two kinetic profiles, 0x1 and 0112 are their standard deviations, and (7sz is the covariance of K1 and K2. A kinetic angle of 0° indicates that the kinetic profiles are completely correlated, while an angle of 90° indicates complete independence. Practically, as the kinetic angle decreases, the degree of kinetic differentiation decreases and so does the amount of kinetic information that can be used to differentiate the analytes. The calculation of figures of merit for multivariate data has been described”. Using these methods, which state that the net analyte signal for the kth analyte in a mixture can be calculated as the portion of the data orthogonal to the data due to all of the other components of the mixture, net analyte signal and 52 selectivity can be easily calculated from the mixture data if the pure component response of the components of the mixture are known: v = ( I - XX“ )1: where v is the portion of data vector (pure component response) u that is orthogonal to data matrix X (which consists of the pure component responses of all of the other components), and X+ is the pseudoinverse of X. The norm of v, Ilvll, is the net analyte signal of the analyte. The selectivity, s, is calculated as the ratio of the norms of the orthogonal part of the pure component response and the pure component response, _|lvl| llull The kinetic net analyte signal and kinetic selectivity for each analyte were calculated from the concentration kinetic profile for the analyte and the sum of the concentration kinetic profiles of all components of the mixture. The rate constant ratio is a measure of the similarity of the reaction rates, and is insensitive to factors other than the rate constant which affect the kinetic profile. The kinetic angle is sensitive to factors other than the rate constant; indeed the length of time for which the reaction is observed has a large effect on the kinetic angle. A plot of kinetic angle vs. rate constant ratio (with other factors that contribute to the kinetic angle held constant) is shown in Figure 3-1. 53 50.0 45.0 4 40.0 4 35.0 a le 5’ 30.0 — 25.0 — 20.0 ~ Kinetic An 15.0 _ 10.0 ~ 5.0 ~ 0.0 r T 1 r 0 2 4 6 8 10 Rate Constant Ratio Figure 3-1: Kinetic angle as a function of rate constant ratio. The fraction of the slower reaction observed was held constant at 90%, and the number of data points was held at 100. Neither the rate constant ratio nor the kinetic angle provide information about which analyte is more resolved. The kinetic net analyte signal does provide this information. For the case where the rate constant ratio is varied (by varying the faster rate constant) and the length of time the reaction is observed is held constant (at the time at which the slower reaction has reached 90% completion), the plot of kinetic net analyte signal vs. rate constant ratio is as shown in Figure 3-2. 54 5.0E-06 4.5E-06 ~ 4.0E-06 ~ 3.55-06 - 3.0E-06 i 2.5E-06 _ 2.0E-06 r 1.5E-06 * Kinetic Net Analyte Signa 1.0E-06 a 5.0E-07 4 0.0E+00 1 . , g Rate Constant Ratio Figure 3-2: Kinetic net analyte signal for the slower (squares) and the faster (circles) reactants in a two-component mixture. The fraction of the slower reaction observed was held constant at 90%, and the number of data points was held at 100. More information is available for the slower analyte than the faster analyte. The analytical signal of the slower component is larger than that of the faster component since it reaches a higher equilibrium concentration. When all other variables are equal and the concentration sets shown later in Figure 3-3 are used, the slower analyte can be more accurately determined than the faster analyte because of its higher net analyte signal. 55 3.2.2. Methods for quantifying spectral differences The degree of spectral overlap (or more precisely, the amount of spectral differentiation) between the analytes was quantified by calculating the angle between the two spectra. This angle, referred to in this work as the spectral angle (05), is defined as the arccosine of the correlation between the spectra (represented by their respective molar absorptivities (81) so that the spectral angle is independent of concentration, and so therefore also of time): ass 65 =arccos ———'——2— 0511752 where 81 and 52 are the two spectral profiles, 0'51 and 0'52 are their standard deviations, and 05152 is the covariance of S1 and S2, A spectral angle of 0° indicates that the spectra are completely correlated, while an angle of 90° indicates complete independence. Practically, as the spectral angle decreases, the degree of spectral differentiation decreases and so does the amount of spectral information that can be used to differentiate the analytes. The synthetic spectra used in the simulations are shown with their spectral angles in Figure 3-3. 56 Spectral Angle=77.7 Spectral Angle=40.7 Spectral Angle=20.0 Spectral Angle=13.5 Spectral Angle=10.3 Spectral Angle=8.6 Spectral Angle=7.6 Spectral Angle=0.0 Spectral Channel Spectral Channel Figure 3-3: Synthetic spectra used for simulation studies. Spectral angles range between 77.7 and 0. A spectral net analyte signal can be calculated in a manner similar to the method for calculating the kinetic net analyte signal. It, like the kinetic net analyte signal, has the advantage of proving information that reveals which analyte is more resolved, and so should be more accurately determined. Figure 3-4, the spectral net analyte signal for each analyte is plotted vs. the spectral analytes between the analytes. The dashed line corresponds to the analyte whose spectrum is depicted with a dashed line in Figure 3-3. 57 9.0E+05 8.0E+05 — 7.0E+05 * 6.0E+05 a 5.0E-1-05 r 4.0E+05 - 3.0E-1-05 a Net Analyte Signal 2.0E+05 4 1.0E-1-05 r ODE-+00 I 1 1 1 O 20 40 60 80 1 00 Spectral Angle Figure 3-4: Spectral net analyte signal as a function of spectral angle. The spectra used to generate this plot are shown in Figure 3-3. The more strongly absorbing analyte (depicted with the dashed line) has a higher net analyte signal. As the spectra become more similar, the difference between the net analyte signals decreases as well. We would thus expect the more strongly absorbing analyte to be determined with greater accuracy. 3.3. EFFECT OF EXPERIMENTAL VARIABLES Data were generated for a collection of 12 standard calibration mixtures and four unknown mixtures. The arrangement of these mixtures in the concentration space of the two analytes is shown below. In all cases, the simulated data were 58 generated using 30 spectral channels (wavelengths); in most cases 100 time points were generated (unless the number of time points was being varied). 5 T T I I I 1' T 1 fir 4.5 - - A 4 ~ 111 a . 2 :1. c 3.5 ~ 0 r .2 on g 3 - 111 111 111 111 . c m g 4 O 2 5 O O 0 g 2 - 111 a a a -l at c g 1.5 r- o _ .2 (D 1 b 111 111 - 0.5 '- - O I 1 1 l 1 l 1 L 1 o 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 Fast Analyte Concentration (11M) Figure 3-5: Concentrations of calibration (stars) and unknown (open circles) samples. These data were used to generate calibration models with each of the multivariate calibration techniques being investigated. These models were applied to data from the four unknown mixtures, and the error of prediction for each analyte was calculated as the relative standard error of prediction (RSEP) 59 % RSEP = i=1 .100 where C, is the true concentration of the analyte in sample i, C,- is the predicted concentration of the analyte in sample i, and h is the number of samples. 3.3.1. Effect of kinetic and spectral angles The kinetic resolution of the analytes was varied by changing the rate constant of the slower analyte and by varying the neamess to completion of the slower reacting analyte at the end of data collection (fraction of the reaction observed). These two experimental parameters were found to influence the accuracy of the determinations through their influence on the kinetic angle, and so on the amount of kinetic differentiation observed between the analytes. In all cases, the slower analyte was more accurately determined. Examining the kinetic net analyte signal (Figure 3-2) reveals that this is expected. The slower reacting analyte was also the more strongly absorbing analyte, and so it has a higher spectral net analyte signal. For all cases studied, principal component regression and partial least squares regression produced predicted concentrations of similar accuracy. Determinations using continuum regression were slightly more accurate. In every case, multiple linear regression produced inferior predictions. PARAFAC generally produced poor predictions, and multiway PLS (nPLS) in most cases produced the most accurate predictions. The following plots illustrate the accuracy 60 of the predicted concentrations produced by each algorithm (expressed as the %RSEP) as a function of kinetic and spectral angles. Figures 3-6 through 3-9 shows the effect of varying kinetic and spectral angles on the accuracy of a PLS, PCR, CR, and nPLS prediction, respectively. PLS, PCR and CR all behave similarly. The accuracy of the nPLS prediction is generally higher, but has the same dependence on the kinetic and spectral angles as do the PLS, PCR and CR determinations. The accuracy with which the faster analyte can be predicted decreases as the spectral angle decreases. The slower analyte is mostly unaffected by a decrease in spectral angle. At any given kinetic angle, the slower analyte has a larger kinetic net analyte signal (Figure 3-2) and so its prediction is less affected by a decrease in spectral differentiation because it is can rely more on kinetic information than can the prediction of the faster analyte. Changes in the kinetic angle have little effect at all but the lowest spectral angles, revealing that the regression algorithms rely more heavily on spectral than kinetic differences. At very low spectral angles (<10°), the amount of spectral information available becomes negligibly small, and the prediction of both analytes depend on the kinetic information. In this case decreasing kinetic angle produces marked increases in the error of both predictions. The error of the prediction of the faster analyte increases more quickly since less kinetic information about it is available. At very low spectral and kinetic angles an accurate determination can not be performed. 61 Fast Analyte Slow Analyte %RSEP (PLS) .~ %RSEP (PLS) M 0| 0 .5 (II N . I I l J SpectralAngIo 4° Spectrum KIneticAflab Figure 3-6: Relative standard error of a PLS prediction as a function of the kinetic and spectral angles. Fast Analyte Slow Analyte %RSEP (PCR) 1" %RSEP (PCR) 30 ‘0 Spectral Angle 4° Spectral Angle Kinetic Angle Kinetic Angle Figure 3-7: Relative standard error of a PCR prediction as a function of the kinetic and spectral angles. 62 Fast Analyte Slow Analyte %RSEP (CR) 1" 4° Spectral Angle Kinetic Angle Kinetic Angle Figure 3-8: Relative standard error of a CR prediction as a function of the kinetic. and snectral angles. Fest Analyte Slow Analyte 4.5 \ ' 3. 9 9 a! 11 5 5 n. n. 2. in ur in m a: a: a? a9 1.5 \ ' 0.5 4 ~ Spectral Angle 4° Spectral Angle Kinetic Angle Kinetic Angle Figure 3-9: Relative standard error of an nPLS prediction as a function of the kinetic and spectral angles. 63 The effects of varying kinetic and spectral angles on an MLR determination are shown in Figure 3-10. The same general trends observed for PLS, PCR, CR, and nPLS can be seen here as well, though the magnitude of the errors is much larger. Feet Analyte Slow Analyte xnssp (MLR) %nsep (MLR) KineticAnge Figure 3-10: Relative standard error of an MLR prediction as a function of the kinetic and spectral angles. Figure 3-11 shows the effect of varying kinetic and spectral angles on the accuracy of a PARAFAC determination. At kinetic or spectral angles of less than 10° neither analyte can be accurately determined, though when both angles are high the accuracy of the predictions rivals that of CR and nPLS. PARAFAC’s flaw, then, is its need for resolution in both, rather than one, dimension. anwme fimkflm a a g a $25 :2 n. n. Lu Lu U) U) I: a: x a 2 1.5 1 _‘ - in O or 9 or 1 Figure 3-11: Relative standard error of a PARAFAC prediction as a function of the kinetic and spectral angles. In the following sections, the results shown will be those produced by continuum regression and multiway partial least squares regression. 3.3.1.1. Contributors to kinetic angle The effect of decreasing kinetic information is not evenly distributed over the two analytes. As already mentioned and seen in Figures 3-6 to 3-11, the 65 prediction of the faster-reacting of the two components is affected to a much larger degree than is the prediction of the slower-reacting analyte. Using CR or nPLS (the techniques to which it can be assumed that most assertions in the next several sections refer) the slower-reacting analyte can be successfully determined in a two-component mixture if either angle is greater than 8°; the faster-reacting component requires that at least one angle be 15° for a determination to succeed. The faster component is also more sensitive to small spectral angles than is the faster component. The fraction of the slowest reaction observed (defined as the fraction of the slower analyte molecules that have reacted at the time observation of the reaction ceases) was changed by varying the length of time over which the reaction was monitored. It was found that kinetic angle is a function of both the ratio of the rate constants and the fraction of the slower reaction observed. 66 Fraction Reaction Observed "8‘9 00mm “360 Figure 3-12: Kinetic angle as a function of rate constant ratio and fraction of the slower reaction observed. The relationship between kinetic angle and these two parameters is shown in Figure 3-12. Generally, kinetic angle increases as either the fraction of the slower reaction observed increases or the ratio of the rate constants increases. The effect of increasing rate constant ratio at a constant 90% reaction observed was shown in Figure 3-1. The same relationship observed there can be seen in Figure 3-12. At low fractions of the reaction observed, the relationship is much the same, though the total change is kinetic angle is smaller. At any constant rate constant ratio, an increase in the fraction of the slower reaction observed results in a fairly linear increase in kinetic angle. 67 3.3.2. Effect of number of data points acquired (time points) The number of time points acquired was varied while the time for which data was taken was held constant (i.e., the rate at which spectra were acquired was varied). In general it was found that varying the number of spectra between 10 and 100 produced no appreciable change in the accuracy of the determination. 3.3.3. Effect of instrumental noise Heteroschedastic (i.e., not uniform across the spectrum) instrumental noise proportional to absorbance at each wavelength was added to all simulated data. In most cases, 1% noise was added; in a few cases this value was varied, and the effect of varying levels of noise on the accuracy of kinetic-spectrophotometric determinations was explored. It was found that as the instrumental noise was varied with kinetic angle (at a low spectral angle) the accuracy of the determination using the multivariate calibration techniques decreased uniformly at all kinetic angles as the noise level increased (see Figure 3-13). 68 Fast Analyte Slow Analyte %RSEP (on) %RSEP (on) Figure 3-13: Relative standard error of a CR prediction as a function of the kinetic angle and instrumental noise. Determinations using multiway PLS were more tolerant of instrumental noise at high kinetic angles as can be seen from Figure 3-14 below. 69 Fast Analyte Slow Analyte %RSEP (nPLS) %nsep (nPLS) 4o lfinetic Angle Figure 3-14: Relative standard error of an nPLS prediction as a function of the kinetic angle and instrumental noise. When the level of the instrumental noise was varied concurrently with the spectral angle (at a constant low kinetic angle), it was found that the effect of the instrumental noise was less pronounced at high spectral angles. Figure 3-15 shows this effect. 70 Fast Analyte Slow Analyte E? E‘ e 9 a a 3: (n a: a: a! 32 Spectral Angle Figure 3-15: Relative standard error of a CR prediction as a function of the spectral angle and instrumental noise. 71 This same phenomenon was observed with nPLS, though the magnitude of the errors was lower. One can see from Figure 3-16 that errors are somewhat lower than from Figure 3-15. Fast Analyte Slow Analyte %nsep (nPLS) %nsep (nPLS) 4O Spectral Angle Spectral Angle % Noise 96 Noise Figure 3-16: Relative standard error of an nPLS prediction as a function of the spectral angle and instrumental noise. It is apparent that instrumental noise (as defined and added to the data) has more influence on the spectral differentiation between the analytes than on the kinetic differentiation, i.e., noise tends to blur spectral differences, especially when the spectra are similar. Techniques that use unfolded data are unable to ignore the spectral data and simply focus on the kinetic data; they are thus 72 unaffected by increased kinetic differentiation. Multiway PLS is able to more exclusively use the kinetic data and so is able to take advantage of increasing angles. At constant kinetic angles, all techniques perform better with high noise levels if the spectral angle is large, though multiway PLS is (as usual) more accurate. . 3.3.4. Effect of rate constant fluctuations Rate constant fluctuations during the course of the reaction (which could be caused by temperature fluctuations or other external perturbations) were simulated by allowing the rate constant to vary with a Gaussian distribution centered on the true value and having a standard deviation proportional to a percentage of the rate constant. Large rate constant fluctuations produce acceptable errors if either the spectral or kinetic angles are moderately large. Indeed, at spectral angles of 20° or larger, rate constant fluctuations up to 15% had no noticeable effects, as can be seen from Figures 3-17 and 3-18. 73 Fast Analyte Slow Analyte 10 ~ - E E 9, e n. o. m u: w w c: at 32 a2 5 4o Kinetic Angle ‘36 Rate Constant Fluctuation % Rate Constant Fluctuation Figure 3-17: Relative standard error of a CR prediction as a function of the kinetic angle and rate constant fluctuation. 74 Fast Analyte Slow Analyte 1 5 \ 15 a 10 q ' i E E 9, 9, LL a. tu Lu «n 0: cc 1: a? a! 5 Spectral Angle Spectral Angle 96 Rate Constant Fluctuation % Rate Constant Fluctuation Figure 3-18: Relative standard error of an nPLS prediction as a function of the kinetic angle and rate constant fluctuation. From comparing Figures 3-13 through 3-16 with Figures 3—17 and 3-18, we can see that the addition of instrumental noise (as defined above) has a much larger effect on the accuracy of determinations than do rate constant fluctuations. For this reason, it was decided to focus more effort on decreasing the instrumental noise level of the experimental system than on connolling factors that could lead to rate constant fluctuations. 75 3.4. EXPLORATION OF EFFECT OF EXPERIMENTAL VARIABLES ON SIMULATIONS OF THE GA(III) -- NI(II) SPECTRAL SYSTEM For these studies, experimentally determined spectra (Figure 3-19) of 4-(2- pyridylazo)-resorcinol (PAR) and its complexes with gallium (III) and nickel (II) were used to generate the simulated data. The procedure by which these spectra were acquired is given in chapter four. 2.5- ..... , .0 .0 o 1.5— ‘.".. -4 Absorptivity (M cm)‘1 Wavelength (nm) Figure 3-19: Absorption spectra of PAR and its Ni(II) and Ga(III) complexes used for simulation studies. 76 The spectral angle between the analytes is 6.3; the spectral net analyte signal for the Ni(H)-PAR complex is 26% higher than that for the Ga(IH)-PAR complex. This indicates that it should be much more accurately predicted. Ni(II) is the faster reacting analyte, and so always has a lower kinetic net analyte signal. In practice, the much higher spectral net analyte signal overwhelms the slightly lower kinetic net analyte signal. It is thus expected that Ni(II) should be determined more accurately than Ga(III). Using the measured spectra, the remaining experimental parameters were varied. The simulations were carried out in the same manner as the previously described studies. 3.4.1. Effect of Kinetic Angle In this highly spectrally overlapped system, it is not unexpected that the kinetic angle should affect the accuracy of kinetic-spectrophotometric determinations (see Figures 3-6 through 3-11). The multivariate calibration techniques are adversely affected by factors that lower the kinetic angle. This is seen in Figure 3-20 which shows that errors below 5% can be achieved for the faster (less spectrally overlapped) analyte under most kinetic conditions, i.e., with all but the lowest kinetic angles (or as shown here, with all combinations of rate constant ratios and fractions of the slower reaction observed save those with low (<3) rate constant ratios and less than 40% of the slower reaction observed). The slower(more spectrally overlapped) analyte is less accurately determined; errors 77 near 5% can be determined only when at least 50% of the slower reaction is observed. Fast Analyte Slow Analyte ‘0 2°\ . .............. E E 0- n- ........... u’ i.“ m (D m m 32 32 5 1o 0 2 Ram ....................... conmm giant Ratio 6 0.8 05 0.4 0.2 s 0.5 03 0.4 0.2 Fraction Reaction Fraction Reaction Observed Observed Figure 3-20: Relative standard error of a CR prediction as a function of the rate constant ratio and the fraction of the slower reaction observed. The Ga(III)/Ni(H) spectra shown in Figure 3-19 were used. The spectral angle was 6.3. Multiway PLS is more able to use the limited kinetic and spectral information available in this system. As such, the accuracy of its predictions is largely unaffected by the parameters that affect the kinetic angle. Looking back at Figure 3—1 1, it can be seen that a spectral angle of ~6, the kinetic angle has little effect. Figure 3-21 shows that neither the rate constant ratio nor the fraction of the slower 78 reaction observed had an appreciable effect on the prediction of either analyte. The faster (less spectrally overlapped) analyte is predicted with greater accuracy. Fast Analyte Slow Analyte _. o 20“; %RSEP (nPLS) %RSEP (nPLS) U! .a O 1 0 Rate 2 Rate ...................... Constant Constant “““ Ratio Ratio . - 6 0'8 05 0.4 0.2 6 0‘8 05 0.4 0.2 Fraction Reaction Fraction Reaction Observed Observed Figure 3-21: Relative standard error of a CR prediction as a function of the rate constant ratio and the fraction of the slower reaction observed. The Ga(III)/Ni(H) spectra shown in Figure 3-19 were used. The spectral angle was 6.3. 79 3.5. (1) (2) (3) REFERENCES Faber, K.; Lorber, A.; Kowalski, B. R. "Analytical figures of merit for tensorial calibration" J. Chemometr. 1997, 11, 419-461. Lorber, A.; Faber, K.; Kowalski, B. R. "Net analyte signal calculation in multivariate calibration" Anal. Chem. 1997, 69, 1620-1626. Messick, N. J.; Kalivas, J. H.; Lang, P. M. "Selectivity and related measures for nth order data" Anal. Chem. 1996, 68, 1572-1579. 80 CHAPTER 4 DETERMINATION OF GALLIUM (III) AND NICKEL (II) No man ’5 knowledge here can go beyond his experience. «John A. Locke The determination of nickel (II) and gallium (IH) using their reaction with 4-(2-Pyridylazo)-resorcinol (PAR) was carried out under a variety of experimental conditions. Kinetic, spectrophotometric, and kinetic-spectrophotometric determinations were performed. Several of chemometric methods were used in the determinations. PAR is a highly sensitive photometric reagent that reacts with a variety of metal ionsl. It is a triprotic weak acid whose dissociation progresses as1: H3L“ —,£_='—>H,L—.i’—+m; ;EE—HLZ' pde pH 3~5.5 pH 6~12.5 pH>12.5 Q—N =N—IQ-OH 0 Figure 4-1: Acid dissociation of PAR In all of the work described in this document, PAR exists in its singly protonated form. PAR complexes with metal ions as either a bi- or tridentate ligand. The two 81 metal ions whose determination is discussed in this chapter, Ga(III) and Ni(II), both form complexes where two PAR molecules react with one metal ionl. Niz*+2 PAR -,2 Ni PAR, Ga3+ +2 PAR -‘-—-‘ Ga PAR2 Gallium (III) and nickel (II) were chosen for several reasons. Both react with PAR, and both are soluble in the buffer solutions used in these studies. They also present a highly challenging system. The rates at which they react with PAR are very close together, and the absorption spectra of their reaction products with PAR are quite similar. The use of PAR as a spectrophotometric reagent for the determination of gallium and especially of nickel has been reported in the literature”. Other workers have described the determination of nickel or gallium using a chromatographic separation and either a pre-concentration step involving PAR or a post-column reaction with PAR10'13. There have been a few reports of kinetic determinations of nickel or gallium using PAR as a reagent14’l5. 4.1. EXPERIMENTAL 4.1.1. Solution Preparation. All solutions were prepared with distilled water and reagent grade chemicals. Buffers of pH 8.5 and 7.0 were prepared from sodium borate and sodium phosphate, respectively, and adjusted with nitric acid. All working solutions were prepared in one of these buffers. 82 The equilibrium studies were performed using a set of three calibration sets. The single component calibration and unknown solution sets contained six and four solutions, respectively, and are described by Tables 4-1 and 4-2. Table 4-1 Single component calibration and unknown sets for the equilibrium determination of Ni(II) Ni(II) concentration (uM) Solution # Calibration Set Unknown Set 1 0.06 2.00 2 1.60 4.00 3 2.60 4 3.60 5 4.60 6 5.60 Table 4-2 Single component calibration and unknown sets for the eguilibrium determination of Ga(III) Ga(III) concentration (1.1M) Solution # Calibration Set Unknown Set 1 1.60 2.00 2 2.60 4.00 3 3.60 4 4.60 5 5.60 6 6.60 The equilibrium multicomponent determinations were performed using a solution set containing eight calibration solutions and two unknown solutions. These samples were reacted with 1mM PAR at a pH of 7.0. The position of these samples in concentration space is shown in Figure 4-2. 83 5 I If r r T I I I I 4.5 L * 4 4 ' r A are 33.5 » . § 3 - 4 Z I.— o o c 2.5 .- * 'i .9 iii 5 «E 2 8 o S 1.5 r- * * - O 1 - _ 0.5 P . 0 1 r l i l l i- t r 1 i ‘1' o 0.5 1 1.5 2 3 3.5 4 4.5 5 2.5 Concentration of Ga(III) (pM) Figure 4-2: Concentrations of calibration (stars) and unknown (open circles) samples used for the two-component equilibrium determinations. For the pH 8.5 kinetics studies, thirteen calibration mixtures and four unknown mixtures were made. These were mixed and reacted in the stopped flow system with a solution of 10’3 M PAR. The calibration and unknown sets used for the pH 8.5 kinetics studies are can be seen in Figure 4-3. 84 60 I I I I I I I T I a a a 50’ i O 3 4 4 340 - ° - 2 s 4 '5 c 30 '- .. .9 :6: o a E c 4 c 2° 0 8 a at 10L J I’ § § 0 l l J l l l P l l 20 25 3O 35 40 45 50 55 60 65 70 Concentration of Ga(lll) (uM) Figure 4-3: Concentrations of calibration (stars) and unknown (open circles) samples used for the ph 8.5 kinetic and kinetic-spectrophotometric determinations of Ga(III) and Ni(H). The pH 7.0 studies were performed using a similar, but slightly different set of solutions. Sixteen calibration and four unknown solutions were mixed and reacted in the stopped flow system with a solution of 10'3 M PAR. The location of the mixtures in the concentration space of the analytes is given in Figure 4-4. 85 20 r I I T T I I I I 18— J 16- ~ A is a a it V %12- a a as a 1 ._ o O c 10* O 'l 3% e a it a l- E 8 o g 6" * fi 4.. * -( 4h— -r 2* . O L 1 l l l 1 l l L 0 5 10 15 20 25 3O 35 4O 45 50 Concentration of Ga(III) (pM) Figure 4-4: Concentrations of calibration (stars) and unknown (open circles) samples used for the ph 7 .0 kinetic and kinetic-spectrophotometric determinations of Ga(III) and Ni(II). 4.1.2. Spectrophotometric (Equilibrium) Data Collection Gallium (III) and Nickel (II) were reacted with PAR in a pH 7.0 phosphate buffer. The reactions were allowed to proceed to equilibrium, and spectra were obtained. Data were collected in two wavelength ranges (500-550nm and 500- 600nm) using a Hitachi U-4001 UV-visible spectrophotometer. A one centimeter quartz cuvctte was employed as a sample holder. A pH 7.0 phosphate buffer was used as a blank. Spectra were collected at 201 wavelengths between 500 and 600 nm, and 101 wavelengths between 500 and 550 nm.. 86 4.1.3. Kinetic-spectrophotometric Data Collection Data were collected using a home-built stopped—flow apparatus interfaced to a thermoelectrically cooled Tracor Northern (Model TN-6123) 512 element intensified diode array (Tracor Northern, Philadelphia, PA) configured to acquire spectra in the 400-800 nm range as described in chapter 2. In the wavelength range of interest (520-560 nm), absorbance was measured at 52 equally-spaced wavelengths. In cases where the entire 520-560 nm spectral region was not used, a subset of the 52 wavelengths was employed. Specifically, in the range 540-560 nm, 26 wavelengths were used. At pH 8.5, kinetic information was obtained by acquiring 26 spectra at a rate of 5.0 scans per second for a total acquisition time of 5.2 seconds. At pH 7.0, 100 spectra were acquired at a rate of 7.575 scans per second over a total acquisition time of 13.07 seconds. 4.1.4. Data processing Time dependent spectra were collected in triplicate and averaged, i.e., each metal ion solution was reacted with PAR three times. The resulting three sets of kinetic-spectrophotometric data were averaged. Data were mean-centered (i.e., the mean of each variable vector was subtracted from each of its elements) before being input to the appropriate algorithms. Multivariate calibration algorithms provided in the PLS_TOOLBOX (Eigenvector Technologies, Manson, WA.) and run in MATLAB were used to perform determinations. 87 4.2. DETERMINATION OF RATE CONSTANTS AND ABSORPTION SPECTRA Experiments were performed to determine pure component spectra and rate constants for the reaction products of the reaction of Ni(H) and Ga(III) with PAR. The pure spectrum of PAR was subtracted from observed absorbance versus time data resulting from the reaction of PAR with a single metal ion. These subtracted data were fit to the equation16’17: A. = A... -(A.. ‘A0)’e-k"t The pseudo-first order rate constant, k’, was calculated from data at several wavelengths, and an average value was computed. The second-order rate constant was calculated from this average k’ and the known excess concentration of PAR as k = k’[PAR] From the fit values of A”, A0, and k’, the initial rate of the reaction was calculated at each wavelength: 6 , A‘ at = k (AW _ A0) The initial rate was plotted versus concentration. The molar absorptivities of the reaction product at each wavelength were computed from the slope of this plot: slope kl These values were then used in the simulations described in other chapters. 88 4.3. RESULTS OF THE DETERMINATION OF GALLIUM (IH) AND NICKEL (II) A spectrophotometric (equilibrium) determination of Ga(III) and Ni(II) was performed, and the ability of the various chemometric algorithms to accurately predict the concentration of the analytes in unknown mixtures was examined. Rate constants and pure component spectra were obtained experimentally for the reactants and products of the reactions of Ni(II) and Ga(III) with PAR. Kinetic- spectrophotometric determinations of Ga(III) and Ni(II) were performed. These determinations were then compared to the spectrophotometric (equilibrium) determination and to a kinetic determination carried out at a single wavelength. 4.3.1. Spectrophotometric (Equilibrium) Determination of Gallium (III) and Nickel (II) Continuum regression and partial least squares regression both produced acceptable results for single-component equilibrium determinations. The relative standard errors of prediction (as defined in chapter two) for each determination are shown in tables 4—3 and 4-4. Table 4-3 Spectrophotometric (equilibrium) determination of N i(II) Method % RSEP % RSEP (500-550 nm) (500-600 nm) CR 1.2 1.2 PCR 20.4 20.4 PLS 1.2 1.2 MLR 1.6 1.9 89 Table 4-4 Spectrophotometric (emiilibrium) determination of Ga(III) Method % RSEP % RSEP (500-550 nm) (500-600 nm) CR 8.6 8.3 PCR 30.6 36.8 PLS 8.6 8.3 MLR 19.6 21.0 The determination of Ni(II) produced relative standard errors of prediction of approximately 1.2% for both wavelength ranges when CR and PLS were used. PCR produced inaccurate predictions for reasons which are not altogether clear. Ga(IH) was less accurately predicted than was Ni(II). Because the molar absorptivity of the Ni-PAR complex is much higher than that of the Ga-PAR complex (see chapter 3), more accurate predictions of N i(II) were expected. The determination of Ga(III) produced slightly lower errors for the larger wavelength range (8.3%) than for the smaller wavelength range (8.6%) when CR and PLS were used. The addition of more wavelengths decreases the error of the Ga(III) determination more than it does the error of the Ni(H) determination because although the additional wavelengths add much more noise than information about the analytes, the algorithms used (CR and PLS) are able to extract the small additional amount of information present in the added wavelengths from the background noise. It is interesting to note that the same effect is not seen when principal component regression or multiple linear regression is employed. It is probable that the ability of continuum regression and partial least squares regression to place lower weights on variables (wavelengths) 90 that vary substantially but are not correlated to the concentrations of the analytes is responsible for their more accurate predictions. The Ga(III) determination is most strongly impacted by this phenomenon, as the lower absorptivity of the Ga-PAR complex results in smaller observed absorbances (and thus less information related to the concentrations being determined). The multicomponent equilibrium determination of Ga(III) and Ni(H) showed significantly higher errors than the single-component determinations. These results are summarized in table 4-5. Table 4-5 Multicomponent spectrophotometric (equilibrium) determination of Ni(II) and Ga(III) 500-550 nm 500-600 nm Method % RSEP % RSEP % RSEP % RSEP (Ga) (Ni) (Ga) (Ni) CR 26 6 21 6 PCR 22 6 21 8 PLS 21 6 21 7 MLR 23 4 24 5 As expected, Ni(II) was determined with greater accuracy than was Ga(III). The spectra are highly overlapped, and so there is little information present that can be used to differentiate the two analytes. Thus, it is not surprising that the additional information that is present in the larger wavelength range results in lowered error for determinations using all algorithms. Figures 4-5 and 4-6 are plots of the predicted vs. actual concentrations. 91 5.5 I I 1 I I I I Predicted Concentration of Ga(III) ( pM) l l 0.5 J— 1 1 1 1 1 1.5 2 2.5 3 3.5 4 4.5 5 Actual Concentration of Ga(III) (11M) Figure 4-5: Plot showing the predicted vs. actual concentrations of Ga(III) for a two-component equilibrium determination. The circles show the predicted concentrations. The solid line has a slope of unity and represents a prediction with no error. The dashed lines show i10% error tolerances. 92 N on b 01 O) r r I 1 r 1 1 1 1 Predicted Concentration of Ni(ll) ( 11M) _a I l O A 41 l l l o 1 2 a 4 s 6 Actual Concentration of Ni(ll) (11M) Figure 4-6: Plot showing the predicted vs. actual concentrations of Ni(II) for a two-component equilibrium determination. The circles show the predicted concentrations. The solid line has a slope of unity and represents a prediction with no error. The dashed lines show i10% error tolerances. 4.3.2. Kinetic Determination of Gallium (III) and Nickel (11) Using the data collected during the kinetic-spectrophotometric determinations, a single-wavelength (550 nm) kinetic determination of Ga(III) and Ni(II) was performed at both pH 7.0 and pH 8.5. The results of these determinations are summarized in table 4-6. 93 Table 4-6 Multicomponent kinetic determination of Ni(II) and Ga(III) pH=7.0 pH=8.5 Method % RSEP % RSEP % RSEP % RSEP (GA) (Ni) (Ga) (Ni) CR 35 26 10 6 PCR 30 20 10 6 PLS 35 19 10 4 MLR 47 31 12 9 The results at pH 8.5 (where a good deal of kinetic differentiation exists between the analytes) are quite comparable with those from the kinetic- spectrophotometric determination (table 4-7). There is little spectral differentiation present in this system, and so the addition of more wavelengths adds little additional information about the analytes. At pH 7 .0 (where much less kinetic differentiation exists) the kinetic determination is clearly inferior to the kinetic- spectrophotometric determination. The additional differentiation between the analytes afforded by the spectral information is, in this case, necessary. 4.3.3. Kinetic-spectrophotometric determination of Gallium(III) and Nickel(II) Two different pH values and two wavelength ranges were used to determine Ga(III) and Ni(II). The results of these determinations are presented in table 4-7. As can be seen, errors of prediction at pH 8.5 were generally half those at pH 7 .0. This can be rationalized in several ways. First, the kinetic angle at pH 8.5 (40.1) is much larger than the angle at pH 7.0 (12.6). This represents a large 94 increase in the amount of kinetic differentiation between the analytes. The spectral angles at pH 8.5 and pH 7.0 are similar, but since both are small (~63), kinetic differentiation is the major source of selectivity; it is thus expected that a larger kinetic angle will result in a smaller error of prediction. Indeed, this is what is observed. Table 4-7 Multicomponent kinetic-spectrophotometric determination of N 101) and Ga(III) 540-560 nm 520-560 nm pH Method % RSEP % RSEP % RSEP % RSEP (Ga) (Ni) (Ga) (NiL 7.0 CR 19 17 17 18 PCR 18 18 20 19 PLS 20 17 19 17 MLR 50 12 11 20 nPLS 21 15 29 16 8.5 CR 11 6 14 8 PCR 9 5 l 1 6 PLS 9 6 14 6 MLR 31 5 29 8 nPLS 9 5 ll 6 The results of these determinations can be compared to the results of the equilibrium spectrophotometric determination described previously (table 4-5). The results of the kinetic-spectrophotometric determination performed at pH 8.5 show a marked improvement over the equilibrium results for all techniques except multiple linear regression. In comparing the multivariate calibration techniques, continuum regression, partial least squares regression, principal component regression and multiway partial least squares regression all produce predictions of similar accuracy. 95 Multiple linear regression sometimes produced slightly more accurate predictions, but, more often, was clearly inferior. For the type of data studied, continuum regression and principal component regression proved the most stable of the multivariate calibration techniques. While not always the best choice, they were rarely worse than any others and often significantly better. Multiway PLS (nPLS) was often superior to the one way techniques, but also performed poorly on occasion. The reasons for these failures are not clear. The spectral region between 540-560 nm was found to contain the majority of the spectral differentiation between the analytes. Increasing the spectral window greatly increased the data processing time, but did not appreciably affect the error of prediction. The results are similar to what is expected based on the simulation studies done using the Ga(III) and Ni(H) spectra and on the relative kinetic and spectral net analyte signals (as discussed in chapter 3). Ni(II) was more accurately predicted, as expected. 96 4.4. (D (D G) M) (5) (O 0) (& (% REFERENCES Shibata, S., "2-Pyridylazo compounds in analytical chemistry", in Chelates in Analytical Chemistry; Flaschka, H. A., Barnard, A. J ., Jr, Eds; Marcel Dekker: New York, 1972; Vol. 4. Bobrowska Grzesik, E.; Grossman, A. M. "Derivative spectrophotometry in the determination of metal ions with 4-(pyridyl-2-azo) resorcinol (PAR)" F resenius J. Anal. Chem. 1996, 354, 498-502. Cladera, A.; Gomez, E.; Estela, J. M.; Cerda, V.; Cerda, J. L. "Computer method for the simultaneous kinetic determination of compounds in mixtures based on the use of diode-array spectrophotometry" Anal. Chim. Acta 1993, 272, 339-344. Gomez, 13.; Estela, J. M.; Cerda, V.; Blanco, M. "Simultaneous spectrophotometric determination of metal-ions with 4-(Pyridyl-2- Azo)Resorcinol (PAR)" F resenius J. Anal. Chem. 1992, 342, 318-321. Kolomiets, L. L.; Pilipenko, L. A.; thud, I. M.; Panfilova, I. P. "Application of derivative spectrophotometry to the selective determination of nickel, cobalt, copper, and iron(III) with 4- (2-pyridylazo)resorcinol in binary mixtures" J. Anal. Chem. 1999, 54, 28-30. Ni, Y. N. "Trace-metal determinations by spectrophotometry with a double chromogenic system and a chemometric approach" Anal. Chim. Acta 1993, 284, 199-205. Ridder, C.; Norgaard, L. "Simultaneous determination of cobalt and nickel by flow- injection analysis and partial least-squares regression with outlier detection" Chemometrics Intell. Lab. Syst. 1992, 14, 297-303. Mori, I.; Kawakatsu, T.; Fujita, Y.; Matsuo, T. "Selective spectrophotometric determination of gallium(III) with 2-(5-bromo-2- pyridylazo)-5-diethylaminophenol in the presence of sodium dodecylsulfate and Brij 35" Anal. Lett. 1999, 32, 613-622. Taljaard, R. E.; van Staden, J. F. "Simultaneous determination of cobalt(II) and Ni(II) in water and soil samples with sequential injection analysis" Anal. Chim. Acta 1998, 366, 177-186. 97 (10) (11) (12) (13) (14) (15) (16) (17) Khalaf, K. D.; MoralesRubio, A.; DelaGuardia, M.; Garcia, J. M.; Jimenez, F.; Arias, J. J. "Simultaneous kinetic determination of carbamate pesticides after derivatization with p-aminophenol by using partial least squares" Microchem J. 1996, 53, 461-471. Chakrapani, G.; Murty, D. S. R.; Mohanta, P. L.; Rangaswamy, R. "Sorption of PAR-metal complexes on activated carbon as a rapid preconcentration method for the determination of Cu, Co, Cd, Cr, Ni, Pb and V in ground water" Journal of Geochemical Exploration 1998, 63 , 145- 152. Karve, M. A.; Khopkar, S. M. "Liquid-liquid-extraction of gallium with high-molecular-weight amines from ascorbate solutions" Chem. Anal. 1993, 38, 469-476. Ming, X. Y.; Wu, Y. H.; Schwedt, G. "HPLC analysis of V, Co, Fe, and Ni By 4-(2-Pyridylazo)- Resorcinol, PAR, and H202 and studies on complex properties influencing retention" F resenius J. Anal. Chem. 1992, 342, 556- 559. Arruda, M. A. Z.; Zagatto, E. A. G.; Maniasso, N. "Kinetic determination of cobalt and nickel byflow-injection spectrophotometry" Anal. Chim. Acta 1993, 283, 476-480. Blanco, M.; Coello, J.; Iturriaga, H.; Maspoch, S.; Riba, J.; Rovira, E. "Kinetic spectrophotometric determination of Ga(III)-Al(III) mixtures by stopped-flow injection-analysis using principal component regression" Talanta 1993, 40, 261-267. Moore, J. W.; Pearson, R. G. Kinetics and Mechanisms ; Wiley: New York, 1991. Mieling, G. E.; Pardue, H. L. "Kinetic method that is insensitive to variables affecting rate constants" Anal. Chem. 1978, 50, 1611-1618. 98 CHAPTER 5 KINETIC-SPECTROPHOTOMETRIC DETERMINATIONS IN SYSTEMS WITH NONLINEAR KINETICS When you can measure what you are speaking about, and express it in numbers, you know something about it. But when you cannot- your knowledge is of a meagre and unsatisfactory kind. «Lord Kelvin Most kinetic determinations are carried out under conditions such that the kinetics of the reaction are pseudo first order in the analyte. In cases where multiple analytes are determined, conditions are usually arranged such that the analytes each react according to pseudo first order kinetics: A+R -—) PA B+R —-> PB g:— = —kAAR = -k,’,A k; = k AR 5%? = —kBBR = -k,’,B k; = kBR A7 = Ace-k}! B, = B0 e4?" Under these conditions, the concentration of the analytes (and of their products) at any time t is linearly related to the initial concentration of the analytes. The 99 condition that must be met is that the concentration of the reagent be sufficiently large that it effectively remains constant over the course of the reaction. Several factors can cause kinetic non-linearity. Most obviously, a low reagent concentration invalidates the assumptions necessary for pseudo first order kinetics. In this case, the concentration of the analyte at any given time is no longer linear with the initial concentration. For the case of a single analyte reacting with the reagent: A+R —> PA 1A- : —kAAR dt A, _ AOR, eprokAI-Roh! R0 Here the concentration of the analyte at time t is a complex nonlinear function of the initial concentration of the analyte and reagent. When two or more analytes are present, a discreet solution for the concentration of an analyte at any time can not be written; rather, the set of differential equations %=—k,-A.R %z—kB-B-R dR E=-(kr'A°R)-(ks-B°R) must be solved numerically. Some other conditions also result in nonlinear kinetics. Perhaps the most common of these is the existence of synergistic effects. Here the rate of reaction of 100 one analyte is a function of the concentration of the other, even if the reagent is present in large excess. dA —=—k ~AoR+ B dt A f( ) In all of these cases, most traditional methods for performing kinetic determinations fail. Some work has been done in the area of performing kinetic determinations under conditions where nonlinear kinetics prevail. Modified nonlinear regression algorithms have been used].2 to determine analytes using second and third order reaction schemes. Several workers have used artificial neural networks to process kinetic data collected from reactions with nonlinear kinetics. Blanco and coworkers3 used an artificial neural network to perform a determination of three analytes. Two react with the reagent according to pseudo first order kinetics; the third follows a complex multistep process that is nonlinear. The ANN produced predictions of acceptable accuracy for all three analytes and outperformed both PLS and PCR. In another study4, Blanco’s group used simulated data to perform a detailed study of the use of PLS for nonlinear kinetic data. Here the nonlinearity was added through the addition of a synergistic term (see above equation). In all but the most nonlinear cases studied, the predictions returned by PLS were acceptably accurate. Blanco and coworkers5 have also explored the use of ANN s in situations where the nonlinearity is introduced by having the analytes present in sufficiently 101 high concentrations that the reaction is pseudo first order in reagent. Benzlyarnine and Butlyamine were determined through their reaction with salicylaldehyde. The ANN outperformed PLS and predicted the concentration of the analytes with only 4% error. Ventura et al.6 used an ANN whose inputs were PCA factors to perform kinetic determinations in the presence of synergistic effects and the inherent nonlinearity of the continuous addition of reagent technique. The results of the experiments were quite impressive and the concentrations of the analytes were predicted with good accuracy (errors of prediction of about 5%). In this chapter a standardized way of reporting the degree of kinetic non- linearity present in a system is developed. The results of simulations and experiments performed under conditions of kinetic non-linearity are presented. 5.1. EXPERIMENTAL 5.1.1. Simulations Simulation studies were performed in a manner similar to that described in chapter 3. The degree of nonlinearity was controlled by varying the concentration of the reagent, and was measured as an angle from linearity. The angle from linearity (0,) was calculated by plotting the concentration of the analyte at a fixed time vs. the initial concentration of the analyte for several samples with different concentrations of the two analytes. The angle 0L is related to the correlation between the initial and time-dependent concentrations. 102 0A A HLA =arccos ° ‘ 0A0 "5A. where A0 and Al are the concentrations of analyte A for several samples at times 0 and t, respectively, 0A0 and 0A, are their standard deviations, and O'Aom is the covariance of A0 and A.. An angle from linearity of 0° indicates complete linearity, while an angle of 90° signifies complete nonlinearity. Practically, when the system is linear, the plot is linear; when there is kinetic nonlinearity the concentration of one analyte affects the concentration of the other at the fixed time. The following figure shows this plot for a situation where linear kinetics prevail. 103 m5 53- g a < s 2.5" " n. O-' ° 2 C 0 go 2r- .1 g o *-' E C... 8" -l C o 1.5" o 3 0|- 3 1.- -4 0.5- - 0 1 1 1 1 1 1 1 1 1 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 Initial Analyte Concnetration (11M) Figure 5-1: Example of a linear (pseudo-first-order) kinetic system. This data was generated with a rate constant ratio of 1.7 and a kinetic angle of 10.2. The reagent was in 281 fold excess The angle from linearity was 0.017. Kinetic nonlinearity causes samples with the same initial concentration of an analyte to have different concentrations of the analyte after a fixed time. Figure 5- 2 shows this effect. The effect of the concentration of the second analyte on the kinetics of the first (because the second analyte uses up some reagent, thereby changing the reaction rate for the analyte in question) can be clearly seen. 104 1.8 r T r r r r 1.6- Analyte Concentration At Two Tune Constants (11M) 0.6 *- l 1 l l J l 0.6 0.8 1.2 1.4 1.6 1.8 1 Initial Analyte Concentration (pM) Figure 5-2: Example of a nonlinear kinetic system. This data was generated with a rate constant ratio of 1.7 and a kinetic angle of 10.2. The reagent was in 1.4 fold excess. The angle from linearity was 6.449. 5.1.2. Solution Preparation Solutions of Ni(II) and Ga(III) were made in a pH 8.5 borate buffer as described in chapter 4. Solutions of PAR at concentrations of 1 mM and 40 11M were also made in the same buffer. The solutions prepared were positioned in the concentration space of the analytes as shown in Figure 5-3 105 18 r r r r I r r r n 16“ " * ‘I 14" O 4 A 2 :1. v12l- * * ¥ *- .. § 0 Z *10" O - O c ‘I ‘U * * .9 8 H F 4 E E O 8 C 6” 'I l- .( O U 4— d 2" .l 0 1 1 1 1 1 1 1 1 1 0 5 10 15 20 25 30 35 40 45 50 Concentration of Ga(lll) (11M) Figure 5-3: Concentrations of calibration (stars) and unknown (open circles) samples used for the kinetic-spectrophotometric determinations of Ga(III) and Ni(II) under linear and nonlinear kinetic conditions. 5.1.3. Kinetic-spectrophotometric Data Collection Data were collected using the stopped-flow apparatus and diode array detector as described in chapter 4. One hundred spectra were acquired; the rate at which spectra were acquired was dependent on the concentration of the PAR. For the 1 mM PAR, 100 spectra were acquired over a period of 50 seconds; when the 40 11M PAR solution was used 100 spectra were acquired over a period of 200 seconds. For this study, a wavelength range of 500-550 nm was empirically 106 chosen. Absorbances were measured at 64 equally-spaced wavelengths in this range. 5.1.4. Data processing Time-dependent spectra were collected in triplicate and averaged, i.e., each metal ion solution was reacted with PAR three times. The resulting three sets of kinetic-spectrophotometric data were averaged. Data were mean-centered before being input to the appropriate algorithms. Multivariate calibration algorithms provided in the PLS_TOOLBOX (Eigenvector Technologies, Manson, WA.) and run in MATLAB were used to perform determinations. 5.2. SIMULATION STUDIES INVOLVING SYSTEMS WITH NONLINEAR KINETICS Detailed simulation studies that explored the effect of nonlinear kinetics on the accuracy of predictions generated by an array of chemometric techniques were conducted. Two limiting cases were considered. In the first, the synthetic spectra described in chapter 3 were used to generate data; in these spectra the reagent does not absorb. The second case was one where the reagent does absorb. These used the Ga(III)-PAR / Ni(H)-PAR spectra which were also used in the studies described in chapter 3. 5.2.1. Systems where the reagent does not absorb In the case where the reagent does not absorb, the only effect of kinetic nonlinearity is the type of effect seen in figure 5-2. The slower analyte is slightly more affected by the nonlinearity, as the faster analyte consumes some of the 107 reagent before the slower analyte has reacted to any appreciable extent. This effect is exaggerated at high kinetic angles where the difference between the reaction rates of the analytes is larger and the faster analyte consumes more of the reagent before the slower analyte begins to react. Fast Analyte Slow Analyte I" at l %RSEP (CR) %RSEP (CR) 20 4° Kinetic Angle 4° Kinetic Angle Angle From Unaarity Angle From Linearity Figure 5-4: Relative standard error of a CR prediction as a function of the kinetic angle and angle from linearity. Figure 5-4 shows the effect of kinetic angle and angle from linearity on the accuracy of a CR prediction. Error of prediction increases as the angle from linearity increases and kinetic angle decreases. The concentration of the slower analyte is predicted more accurately; this is expected, as when the determination is 108 limited by kinetic information the slower analyte is favored by the reaction scheme (see chapter 3). The spectral angle is low (8.6), and so the effects of the kinetic angle can easily be seen. The angle from linearity has a much larger effect at high kinetic angle than low ones (for the reasons already discussed). 3.5—1.”. u L 96385? (nPLS) to or 1 to 1 i7. 1 - ....... . ..... . ------- ....... . ...... 0.. ._.. 2 1 20 4° Kinetic Angle Angle From Linearity Slow Analyte ‘ 4o Kinetic Angle Angle From Linearity Figure 5-5: Relative standard error of an nPLS prediction as a function of the kinetic angle and angle from linearity Figure 5-5 shows the response of multiway PLS to the same conditions depicted in figure 5-4. Here, it can be seen that the slower analyte is more affected by kinetic non-linearity, especially at high kinetic angles. It can also be seen through comparison with Figure 5-5 that multiway PLS produces more accurate 109 predictions than do the multivariate calibration techniques (of which CR is the best). At a constant (low) kinetic angle the interaction of spectral angle with kinetic nonlinearity can be demonstrated. Fast Analyte Slow Analyte 5 ~ 5 a 4.5 4 4.5 ~ 4 a 4 ~ 3.5 ‘1 3.5 ‘- A 3 ~ A a: a: e 9 $ 2.5 ~ g 2. w to t: a: a2 2 g 39 1.5 ~ 1 1 a 0 5 ~ 0.5 ~ 0 J _ o o 2.5 2 1.5 1 50 2.5 2 1.5 1 ' 50 0.5 Spectral Angle 0.5 Spectral Angle Angle From Linearity Angle From Linearity Figure 5-6: Relative standard error of a CR prediction as a function of the spectral angle and angle from linearity. The kinetic angle for this data was ~10°. In figure 5-6, it can be clearly seen that at high spectral angles kinetic non- linearity has little affect on the accuracy with which multivariate calibration techniques can predict the concentration of the faster analyte. The prediction of the slower analyte is more strongly impacted by the angle from linearity. At low 110 spectral angles, the kinetic information becomes important, and the slower analyte is predicted with greater accuracy. The increase in error of prediction with increasing angle from linearity can be seen for both analytes. %RSEP (nPLS) 33 u to 1 1 Fast Analyte 0-5 Spectral An 9 Ange From Unearity g %RSEP (nPLS) N a: u L l N J 1.5»“1 1 ' 5° 05 SpectralAngle Slow Analyte 4“ fffff ..... ..... ~~~~~~ ‘‘‘‘‘‘ ........ ........ ......... ....... ..\. Angle From Linearity Figure 5-7: Relative standard error of an nPLS prediction as a function of the spectral angle and angle from linearity. The kinetic angle for this data was ~10°. Figure 5-7 depicts the response of multiway PLS to various spectral angles and angles from linearity. The same general trends described above are visible in these plots, but it is apparent that the errors of prediction are lower. Again, multiway PLS proves superior to traditional multivariate calibration techniques. 11] 5.2.2. Systems where the reagent does absorb When the reagent absorbs in the spectral region of interest the effect of kinetic nonlinearity is complicated. In addition to a nonlinear relationship between initial and fixed—time analyte concentrations, the non-constant reagent concentration contributes a time-dependent rather than a constant spectral background. This effect is becomes more pronounced as the reaction proceeds and the reagent is consumed. This results in less accurate determinations of the slower analyte, especially at high rate constant ratios where the faster reactant has used more of the reagent before the slower analyte has reacted to any appreciable degree. Figure 5-8 shows the results of a CR prediction for a system with the gallium-nickel spectra described in chapter 3. The determination of the faster analyte is not greatly affected by decreasing kinetic angle or reagent excess. The determination of the slower analyte, however, suffers loss of accuracy as the reagent excess decreases, especially at high rate constant ratios (kinetic angles). 112 Fast Analyte Slow Analyte %RSEP (CR) %RSEP (CR) Reagent Excess Figure 5-8: Relative standard error of a CR prediction as a function of the kinetic angle and reagent excess. The gallium/nickel/PAR spectra (with the PAR absorbing) were used to generate the data. The spectral angle for this data was ~11°. Figure 5-9 shows the results of applying nPLS to the same data used for Figure 5- 8. In general, nPLS seems more adversely affected by increasing non-linearity than does CR. 113 Fast Analyte %RSEP (nPLS) %RSEP (nPLS) Figure 5-9: Relative standard error of an nPLS prediction as a function of the kinetic angle and reagent excess. The gallium/nickel/PAR spectra (with the PAR absorbing) were used to generate the data. The spectral angle for this data was ~11°. 5.3. DETERMINATION OF GALLIUM (III) AND NICKEL (11) IN A SYSTEM WITH NONLINEAR KINETICS A determination of Ga(III) and Ni(H) was carried out as described in sections 5.1.2 through 5.1.4. The degree of nonlinearity was determined by varying the concentration of PAR. Two PAR concentrations were used. At the higher concentration (1.0 mM) there is an average reagent excess of 28.1 fold. Here the excess of reagent is calculated by dividing the PAR concentration by the 114 average sum of the analyte concentrations. At the lower PAR concentration (40 M) the PAR is present in 1.2 fold excess. Table 5-1: Kinetic-spectrophotometric determination of Ga(III) and Ni(II) with two different values of reagent excess 1.0 mM PAR 40 11M PAR Method % RSEP % RSEP % RSEP % RSEP (Ga) (Ni) (Ga) (Ni) CR 19 4 28 4 PCR 19 6 34 4 PLS 19 4 43 6 MLR 26 14 56 4 nPLS 21 5 49 4 The results of these determinations are presented in Table 5-1. Inspection of the data reveals that, as is expected based on the results shown in Figures 5-8 and 5-9, the determination of Ni(II) is not affected by the change in PAR concentration. The determination of Ga(III) is more strongly affected by the kinetic nonlinearity present in the data due to the lower PAR concentration. Figures 5-10 and 5-11 show the results of predictions using CR. As in chapter 4, the solid center line in these plots represents a perfect prediction (a slope of unity). The two dashed lines represent 10% error limits. 115 d .e d .e .5 o N & Q m 8 Predicted Concentration of Ni(ll) ( uM) O 45 1.0 115 20 Actual Concentration of Ni(ll) (uM) Figure 5-10: CR determination of Ni(II) by reaction with 1mM PAR (circles) and 40 11M PAR (plus signs). At the PAR concentration of 1mM the average excess is 28 fold; at 40 uM it is 1.2 fold. Figure 5-10 depicts the prediction of nickel. The plot clearly shows that the prediction is largely unaffected by the kinetic nonlinearity. 116 8 8 Predicted Concentration of Ga(lll) ( 1.1M) 23 8 d o o L I l l I l l l 5 1o 15 20 25 so 35 4o 45 so Actual Concentration of Ga(III) (uM) Figure 5-11: CR determination of Ga(III) by reaction with 1mM PAR (circles) and 40 uM PAR (plus signs). At the PAR concentration of 1mM the average excess is 28 fold; at 40 uM it is 1.2 fold. In Figure 5-11 the results of the prediction of Ga(III) can be seen. The overall error is large (28%, see table 5-1), but three of the four unknown samples are predicted well. The fourth, with the highest concentration of Ga(IH), is predicted quite poorly. The PAR is not in excess for this sample, and so the prediction fails. This result is of extreme interest, as it shows that reagent excesses of just a single fold are sufficient for accurate determinations and that only when the reagent is the limiting reactant does the kinetic-spectrophotometric determination fail. This is important, as it reveals that large excesses of extremely 117 high absorptivity reagents in order to ensure pseudo-first order conditions are not necessary. This, then, is perhaps one of the most compelling and powerful arguments for the use of multivariate calibration techniques for kinetic and kinetic-spectrophotometric determinations. 5.4. (1) (2) (3) (4) (5) (6) REFERENCES Schechter, 1.; Schroder, H. "Error-compensated kinetic determinations in systems of mixed lst-order and 2nd-order reactions, without prior knowledge of reaction constants" Anal. Chem. 1992, 64, 325-329. Schechter, I. "Simultaneous determination of mixtures by kinetic-analysis of general-order reactions" Anal. Chem. 1992, 64, 729-737. Blanco, M.; Coello, J.; Iturriaga, H.; Maspoch, S.; Redon, M. "Artificial neural networks for multicomponent kinetic determinations" Anal. Chem. 1995, 67, 4477-4483. Blanco, M.; Coello, J .; Iturriaga, H.; Maspoch, S.; Redon, M. "Partial least- squares regression for multicomponent kinetic determinations in linear and nonlinear-systems" Anal. Chim. Acta 1995, 303, 309-320. Blanco, M.; Coello, J .; Iturriaga, H.; Maspoch, S.; Redon, M.; Villegas, N. "Artificial neural networks and partial least squares regression for pseudo- first-order with respect to the reagent multicomponent kinetic- spectrophotometric determinations" Analyst 1996, 121, 395-400. Ventura, S.; Silva, M.; Perez-Bendito, D.; Hervas, C. "Computational neural networks in conjunction with principal component analysis for resolving highly nonlinear kinetics" J. Chem. Inf. Comput. Sci. 1997, 37, 287-291. 118 CHAPTER 6 DETERMINATION OF Zn(II) AND Cu(II) IN A DRINKING WATER SAMPLE D0, or do not. There is no try. --George Lucas: The Empire Strikes Back (Yoda) The determination of zinc (II) and copper (II) in a standard drinking water sample was carried out using their reaction with PAR. Good accuracy was achieved in spite of the presence of interfering species. Both Zn(II) and Cu(II) react with PAR according to a 1:2 stoichiometrylz Zn“ +2PAR :2 Zn PAR2 Cu2+ +2PAR .-_\ Cu PAR, The determination of zinc and copper in environmental and clinical samples is of some importance? While both metals are essential in small concentrations, they are toxic at higher concentrationsz. The EPA has set limits of 1.3 ppm for copper in drinking water3 and 5 ppm for zinc in drinking water3. The use of PAR as a spectrophotometric reagent for the determination of copper and zinc has been reported in the literature“? Other workers have described the determination of zinc or copper using a chromatographic separation and either a pre-concentration step involving PAR or 119 a post-column reaction with PAR14-23. There have been a few reports of kinetic determinations of copper or zinc using PAR as a reagent2034’25. 6.1. EXPERIMENTAL The sample in which the Zn(II) and Cu(II) were determined was purchased from NSI Solutions, Inc. (Research Triangle Park, NC). It contained a mixture of 11 metal cations in a standard drinking water matrix. These metals were certified to be present in the following concentrations: Table 6-1 Certified concentrations (ppb) of metal cations in a drinking water sample Certified Certified Metal Concentration Concentration (Ppb) 11M Arsenic 113 1.5 Beryllium 5.17 0.6 Cadmium 15.4 0.1 Chromium 103 2.0 Copper 834 13.1 Lead 69.4 0.3 Manganese 222 4.0 Mercury 10.3 0.05 Nickel 246 4.2 Selenium 23.2 0.3 Zinc 1 130 17.3 Of these metals, all but arsenic, beryllium and selenium will react with PARl. 6.1.1. Solution Preparation. All solutions were prepared with distilled water and reagent grade chemicals. A buffer of pH 8.5 was prepared from sodium borate and adjusted with nitric acid. All working solutions were prepared in this buffer, though some concentrated stock solutions were made in water and then diluted with buffer. 120 Twelve calibration mixtures in the micromolar concentration range were prepared from 1 mM stock solutions of the metal nitrates. The Cu(II) concentration was varied between 11 and 17 M; the Zn(H) concentration ranged from 15 to 21 M. These were mixed and reacted in the stopped-flow system with a solution of 100 M PAR in pH 8.5 buffer. A plot of the calibration set in concentration space is given in figure 6-1. 22 1 l r I r I r 21 F 111 a _ 20+ - 19- e a a a - Concentration of Zn(ll) (uM) 16'- - 15 - I a - 14 I l l L l I l 10 11 12 13 14 15 16 17 18 Concentration of Cu(II) (uM) Figure 6-1: Calibration set (stars) used in the determination of Zn(II) and Cu(II). The unknown sample is depicted by an open circle. 121 6.1.2. Kinetic-spectrophotometric Data Collection Data were collected using a home-built stopped-flow apparatus interfaced to a thermoelectrically cooled Tracor Northern (Model TN-6123) 512 element intensified diode array (Tracor Northern, Philadelphia, PA) configured to acquire spectra in the 400-800 nm range as described in chapter 2. In the wavelength range of interest (500-550 nm), absorbance was measured at 64 equally-spaced wavelengths. Kinetic information was obtained by acquiring 22 spectra at a rate of 10.5 spectra per second for a total acquisition time of 2.0 seconds. The spectra and kinetics of the analytes and their reaction products are described in detail later in this chapter. 6.1.3. Data processing Time dependent spectra were collected in quadruplicate and averaged, i.e., each metal ion solution was reacted with PAR four times. The resulting four sets of kinetic-spectrophotometric data were averaged. Data were mean-centered (as described in chapter 3) before being input to the appropriate algorithms. Multivariate calibration algorithms provided in the PLS_TOOLBOX (Eigenvector Technologies, Manson, WA) and run in MATLAB were used to perform determinations. 6.2. DETERMINATION OF ZINC (II) AND COPPER (II) Zn(H) and Cu(II) were determined in both a standard drinking water sample and in a series of synthetic unknown samples. Ni(II) and Mn(II) were the major interfering species in the drinking water unknown and so these were added to the 122 calibration set solutions, and to several of the synthetic unknowns. The calibration samples were all 4.0 uM in Mn(II) and 4.2 11M in Ni(II). The concentrations of the analytes and the levels of the interferents present in each unknown sample are given in Table 6-2. Table 6-2 Concentrations of analytes and interferents in the unknown sam les in uM units Drinking Water Synthetic Unknow Pynthetic Unknowri Synthetic Unknown #1 #2 #3 Cu 13.1 13.1 13.1 13.1 Zn 17.3 17.3 17.3 17.3 Mn 4.0 4.0 13.1 4.0 Ni 4.2 -- -- 4.2 Cd 0.1 -- -- -- Cr 2.0 -- -- -- Pb 0.3 -- -- -- fig 0.05 -- -- -- The results of the determination of Cu(II) and Zn(II) in each of these samples are summarized in Table 6—3. Table 6-3 Results of the determination of Cu(II) and Zn(II) in a series of unknown samples Drinking Water Synthetic Unknow Synthetic Unknow Synthetic Unknown #1 #2 #3 Method % RSEP % RSEP % RSEP % RSEP % RSEP % RSEP % RSEP % RSEP (Cu) (Zn) (Cl!) (Zn) (Cu) (Zn) (Cu) (Zn) PLS 6 6 ll 3 14 2 7 3 PCR 6 6 ll 3 13 2 6 3 MLR l 4 9 l 9 3 9 3 CR 5 1 ll 3 14 2 11 1 Parafac 5 3 21 2 27 2 19 5 nPLS 6 6 11 3 14 2 7 3 The EPA has designated acceptance limits for the determination of Cu(II) and Zn(II) in the drinking water sample. The acceptable range for the determination of copper is $9.8 % error. Acceptable determinations of zinc have 123 errors not larger than i7.98%. All of the chemometric techniques used produced predicted analyte concentrations of acceptable accuracy. Copper was determined with an error of approximately 5%; zinc with an error of 3-6%. Both are well within the EPA guidelines. The presence of the interferents had little effect on the determination. In the second synthetic unknown, the manganese concentration is more than three times its level in the calibration set. The accuracy with which the analytes were predicted is only slightly worse (e.g., 14% vs. 11% error for the CR determination of copper) than it is for the first synthetic unknown. In the third synthetic unknown, which has the interfering species present in the same concentrations as the calibration set, the analytes are determined with better accuracy (an average decrease of 1-3% relative error compared to the first synthetic unknown). The calibration step compensates for the analytical signal due to the interfering species, and so the third unknown is least affected by the interferents. The other two synthetic unknowns are missing Ni(II) (see Table 6-2) , and so are not well described by the calibration set, which does contain nickel. In all cases, zinc is determined more accurately than is copper; the error with which Zn(II) is determined is generally half to one third that of Cu(II) (see Table 6-3). This can be explained by examining the spectral and kinetic contributions to the net analyte signal for each analyte. As a first step, the spectra of the analytes (or specifically, of their reaction products with PAR) can be compared. 124 0.016 I I I I I I I I I 0.014- _ I I 0.012 0.01 - 0.008 - Absorptivity ( uM cm)‘1 0.006 r- 0.004 *- 0 l l l I l l I l l 500 505 510 515 520 525 530 535 540 545 550 Wavelength (nm) Figure 6-2: Spectra of the PAR complexes of Zn(II) and Cu(II). The zinc complex is shown as a solid line, the copper complex as a dotted line. The spectral angle can be computed to be 17.1, a sufficiently high value to suggest that a reasonable determination might be attempted. The spectral net analyte signals reveal that zinc is likely to be more accurately predicted than copper; zinc’s spectral net analyte signal is 5.8 times larger than copper’s. This is true because although the copper complex has a higher molar absorptivity, zinc is present in a high concentration. 125 18 I I I I t r T j l .5 5 f L ............. ................... d d O N r I L 1 Product Concentration ( 11M) 0 0 0.2 0.4 0.6 0.8 1.2 1.4 1.6 1.8 2 T111116 (s) Figure 6-3: Kinetic profiles of the formation of the PAR complexes of Zn(H) and Cu(II). The zinc complex is shown as a solid line, the copper complex as a dotted line. The kinetics of the reactions under the conditions in which the drinking water sample were determined are shown in Figure 6-3. The rate constants for the two reactions are very close; the ratio of the rate constants is 1.1, and the kinetic angle is 2.4. The kinetic data provides slightly more information about the slower reaction of zinc with PAR; zinc’s kinetic net analyte signal is 1.3 times larger than copper’s. Since zinc has both a larger kinetic and spectral net analyte signal, it is not surprising that it is determined with greater accuracy (usually between half and a third the error of the copper determinations—see Table 6-3). 126 6.3. (1) (2) (3) (4) (5) (6) (7) (8) (9) REFERENCES Shibata, S., "2-Pyridylazo compounds in analytical chemistry", in Chelates in Analytical Chemistry; Flaschka, H. A., Barnard, A. J., Jr, Eds; Marcel Dekker: New York, 1972; Vol. 4. Lobinski, R.; Marczenko, Z. Spectrochemical Trace Analysis for Metals and MetalloidsWilson and Wilson’s Comprehensive Analytical Chemistry, ; Elsevier: New York, 1996. Environmental Protection Agency Office of Ground Water and Drinking Water, "Technical Drinking Water and Health Contaminant Fact Sheets" [Online] Available http://www.epa.gov/OGWDW/dwh/t-ioc/zinc.html, 1999 Bobrowska Grzesik, E.; Grossman, A. M. "Derivative spectrophotometry in the determination of metal ions with 4-(pyridyl-2-azo) resorcinol (PAR)" F resenius J. Anal. Chem. 1996, 354, 498-502. Engstrom, E.; Jonebring, 1.; Karlberg, B. "Assessment of a screening method for metals in seawater based on the non-selective reagent 4-(2- pyridylazo)resorcinol (PAR)" Anal. Chim. Acta 1998, 371, 227-234. Fernandez deCordova, M. L.; Molina Diaz, A.; Pascual Reguera, M. I.; Capitan Vallvey, L. F. "Determination of Trace Amounts of Copper With 4-(2- Pyridylazo)Resorcinol By Solid-Phase Spectrophotometry" F resenius J. Anal. Chem. 1994, 349, 722-727. Gomez, E.; Estela, J. M.; Cerda, V.; Blanco, M. "Simultaneous spectrophotometric determination of metal-ions with 4-(Pyl‘idyl-2- Azo)Resorcinol (PAR)" F resenius J. Anal. Chem. 1992, 342, 318-321. Kolomiets, L. L.; Pilipenko, L. A.; thud, I. M.; Panfilova, I. P. "Application of derivative spectrophotometry to the selective determination of nickel, cobalt, copper, and iron(III) with 4- (2-py1idylazo)resorcinol in binary mixtures" J. Anal. Chem. 1999, 54, 28-30. Manouri, O. C.; Papadimas, N. D.; Salta, S. E.; Ragos, G. C.; Demertzis, M. A.; Issopoulos, P. B. "Three approaches to the analysis of zinc(II) in pharmaceutical formulations by means of different spectrometric methods" Farmaco 1998, 53, 563-569. 127 (10) (11) (12) (13) (14) (15) (16) (17) (18) Molina, M. F.; Nechar, M.; Bosque-Sendra, M. "Determination of zinc in environmental samples by solid phase spectrophotometry: Optimization and validation study" Analytical Sciences 1998, 14, 791-797. Ni, Y. N. "Trace-metal determinations by spectrophotometry with a double chromogenic system and a chemometric approach" Anal. Chim. Acta 1993, 284, 199-205. Pollak, M.; Kuban, V. "Comparison of spectrophotometric methods of determination of Zinc(II) in biological-material and study of its complex- formation reactions with 4-(2-Pyridylazo)Resorcinol" Collection of Czechoslovak Chemical Communications 1979, 44, 725-741. Ren, S. X.; Gao, L. "Simultaneous spectrophotometric determination of copper(II), lead(II) and cadmium(II)" Journal of Automatic Chemistry 1995, 17,115-118. Al-Shawi, A. W.; Dahl, R. "The determination of cadmium and six other heavy metals in nitrate phosphate fertilizer solution by ion chromatography" Anal. Chim. Acta 1999, 391, 35-42. Cardellicchio, N.; Dell’Atti, A.; Giandomenico, S.; Di Leo, A.; Cavalli, S. "Determination of transition metals in mineral waters by ion chromatography and spectrophotometric detection" Annali Di Chimica 1998, 88, 819-827. Cardellicchio, N.; Cavalli, S.; Ragone, P.; Riviello, J. M. "New strategies for determination of transition metals by complexation ion-exchange chromatography and post column reaction" J. Chromatogr. A 1999, 847, 251-259. Chakrapani, G.; Murty, D. S. R.; Mohanta, P. L.; Rangaswamy, R. "Sorption of PAR-metal complexes on activated carbon as a rapid preconcentration method for the determination of Cu, Co, Cd, Cr, Ni, Pb and V in ground water" Journal of Geochemical Exploration 1998, 63, 145- 152. de Jesus, D. S.; Casella, R. J.; Ferreira, S. L. C.; Costa, A. C. S.; de Carvalho, M. S.; Santelli, R. E. "Polyurethane foam as a sorbent for continuous flow analysis: Preconcentration and spectrophotometric determination of zinc in biological materials" Anal. Chim. Acta 1998, 366, 263-269. 128 (19) (20) (21) (22) (23) (24) (25) Hardy, S.; Jones, R; Riviello, J. M.; Avdalovic, N. "Construction and investigation of a post-capillary reactor for trace metal analysis by capillary electrophoresis" J. Chromatogr. A 1999, 834, 309-320. Lucy, C. A.; Dinh, H. N. "Kinetics and equilibria of the Zn-EDTA-PAR postcolumn reaction detection system for the determination of alkaline- earth metals" Anal. Chem. 1994, 66, 793-797. Nagaosa, Y.; Tanizaki, M. "Simultaneous determination of zinc(II) and iron(III) in human serum by liquid chromatography using post-column derivatization with 4-(2-pyridylazo)-resorcinol" Journal of Liquid Chromatography & Related Technologies 1997, 20, 2357-2366. Shotyk, W.; Immenhauser Potthast, I. "Determination of Cd, Co, Cu, Fe, Mn, Ni and Zn in coral skeletons by chelation ion chromatography" J. Chromatogr. A 1995, 706, 167-173. Vasconcelos, M. T.; Gomes, C. A. R. "Limitations of ion chromatography with postcolumn reaction for determination of heavy-metals in waters containing strong chelating-agents" J. Chromatogr. A 1995, 696, 227-234. Cladera, A.; Gomez, E.; Estela, J. M.; Cerda, V.; Cerda, J. L. "Computer method for the simultaneous kinetic determination of compounds in mixtures based on the use of diode-array spectrophotometry" Anal. Chim. Acta 1993, 272, 339-344. Taljaard, R. E.; van Staden, J. F. "Simultaneous determination of cobalt(II) and Ni(II) in water and soil samples with sequential injection analysis" Anal. Chim. Acta 1998, 366, 177-186. 129 CHAPTER 7 CONCLUSIONS AND FUTURE PERSPECTIVES Data without generalization is just gossip. «Robert Pirsi g The preceding chapters have described the application of chemometric data processing techniques to kinetic-spectrophotometric data 'under a variety of circumstances. In this chapter, the work as a whole is summarized, and thoughts on future work are presented. 7.1. CONCLUSIONS AND SUMMARY Chapter one presents a fairly detailed overview of the fields of chemometrics and kinetics as applied to analytical chemistry and highlights the application of chemometric techniques to kinetic data. Multivariate calibration algorithms are described in detail; some attention is given to artificial neural networks, multiway algorithms, and other techniques as well. A new data acquisition system was designed and built. This redesign of the existing system involved the fabrication of a new optical path for the stopped-flow apparatus as well as the creation of a new computerized interface for the diode array detection system. The new system was characterized; the results of this characterization and the details of the new designs are found in chapter two. The 130 redesign of the system made possible the collection of the kinetic- spectrophotometric data analyzed in the remaining chapters. In chapter three the results of a series of simulated experiments are described. In these simulation studies the effect of an array of experimental variables on the accuracy of a kinetic-spectrophotometric determination of a two component mixture are explored. Methods for quantifying the amount of kinetic and spectral information present in kinetic-spectrophotometric data were discussed. The kinetic and spectral angles were introduced and shown to be good measures of the quantity of information available in each dimension. Kinetic and spectral net analyte signals were also developed and were shown to be good predictors of the relative accuracy with which analytes can be determined. Simply put, the kinetic and spectral angles allow prediction of the general feasibility of a determination and the dimension (kinetic or spectral) that will be relied upon most heavily. From these data and the net analyte signals it is possible to infer which analyte will be determined most accurately. As an example, the case of gallium and nickel discussed in chapter four can be cited. The kinetic and spectral angles make it clear that spectral information is more heavily relied upon than kinetic in this case. Thus, although gallium has a higher kinetic net analyte signal, nickel’s greater spectral net analyte signal indicates that it will be determined more accurately. This indeed is the result of both the simulated and experimental determinations. 131 The kinetic angle was shown to have several contributing factors. The ratio of the analyte rate constants is the largest contributor, but the fraction of the slower reaction for which data is acquired and the number of spectra acquired also impact the kinetic angle. Chapter four describes the determination of Ga(III) and Ni(II). In the studies discussed, the effect of the kinetic angle was explored experimentally; the determination was carried out at two different values of solution pH where the ratios of the reaction rate constants are different. The expected results were obtained; the determination was more accurate at the pH (8.5) where the kinetic angle was largest. A comparison was drawn between the various chemometric algorithms and it was found that continuum regression generally out-performed the other multivariate calibration techniques, but that partial least squares regression and principal component regression were nearly as good. PARAFAC was found to be useful only in limited circumstances, and multiway PLS showed an ability to compete with and often surpass the best of the multivariate calibration techniques when dealing with multiway data. In chapter five the effect of kinetic non-linearity was examined. Causes of kinetic non-linearity were discussed and recent work in the area of kinetic determinations in nonlinear systems was highlighted. The degree of kinetic non- linearity was measured as an angle from linearity. This angle was shown to be a good predictor of the accuracy with which a determination could be performed. The degree of non-linearity was varied in a series of simulations. In general, it was 132 found that the techniques used were fairly tolerant of nonlinear kinetics, but that at very large angles from linearity the determinations became inaccurate. Experimental studies in which the degree of linearity was varied were also described. In these studies Ga(III) and Ni(II) were determined under two sets of experimental conditions with varying degrees of non-linearity. The results obtained paralleled the simulations and were quite encouraging. Again, the various chemometric algorithms were compared, and similar results to those found in chapter three were obtained. Continuum regression and nPLS proved best suited to handling the nonlinear kinetic data. Chapter six presents the determination of Cu(II) and Zn(II) in a real sample with clinical and environmental relevance. The accuracy of the determination was within the EPA’s acceptance limits for both analytes. The kinetic and spectral angles and net analyte signals were used to explain the relative accuracy with which the analytes were determined. 7 .2. REFLECTIONS ON FUTURE DIRECTIONS All of the work described in this document has focused on the determination of two analytes. The extension to three or more analytes is certainly logical, but also is highly challenging and will present many difficulties. Not least of these will be the need to develop new methods for quantifying the amount of kinetic and spectral information available. Ratios of rate constants and angles between profiles lose meaning when more than two analytes are present. Net analyte signals are better, but are still not capable of the necessary resolution. The 133 net analyte signals reveal the degree to which a profile (kinetic or spectral) is overlapped by other profiles, but do not reveal which specific profiles overlap the profile of interest; e.g., an analyte might be highly spectrally and kinetically overlapped, but by different species. In this situation, the net analyte signals will both be small, but the determination can still be accurately performed since the analyte is not overlapped in both dimensions by the same species. Other problems that will be encountered include the determination of an unoverlapped analyte in the presence of two other highly overlapped analytes, and the economics of scale associated with three-analyte calibration sets. The trend in chemometrics is toward the wider use of multiway algorithms. As they become more popular, their use with multiway data will supercede the use of first order algorithms on unfolded multiway data. This has had and will continue to have an impact on the way in which data are collected and used. The use of multiway algorithms for kinetic-spectrophotometric data has been briefly explored in this work, and it is expected that it will continue to be an active area of study. In 1993 Crouch1 identified seven trends in kinetic methods of analysis. Several of these are relevant to the work discussed in this document. These trends are: 1. Increasing use of “intelligent automation” 2. Growing utilization of multidimensional instrumentation 3. Continuing development of sophisticated data processing techniques 134 4. Additional progress in error-compensation techniques 5. Innovations in multicomponent kinetic procedures 6. Expanding applications of kinetic determinations 7. Enlarging the kinetic approach to include miscellaneous time-dependent responses The work in this thesis is testament to the continuation of trends 2, 3, 4, and 5. As the field of chemometrics continues to advance and to become more widely accessible and accepted, trend three will hold as new (at least to kinetics researchers) chemometric techniques are applied to kinetic data. It is the contention of this author that wider accessibility of and familiarity with chemometric techniques and the proliferation of multidimensional instrumentation will eventually result in an increase in the number of analytical measurements employing kinetic or other time-dependent data. This is in accordance with the sixth and seventh trends listed above. As Crouchl'2 and Mottola3 have pointed out, most analytical methods involve some sort of kinetic or transient response. Often, great pains are taken to ensure that these time-dependent responses are avoided or that they are compensated for in the data processing. Modern multiway chemometric data processing options are able to correctly handle higher order data and can separate and make use of several data dimensions. This implies that time-dependent data can be used in almost all analytical measurements, resulting in increases in selectivity and (in many cases) sensitivity'v3-5. Some of the areas in which the transient data is already being 135 applied include luminescence lifetime spectroscopy, fluorescence lifetime imaging microscopy, sequential injection analysis and transient electroanalytical chemsitry. In all, the future of analytical chemistry appears bright. Greater computing power allows the acquisition of more data in more dimensions and grants the ability to use these data to perform more and better determinations. The field of kinetic methods will no doubt benefit from this trend, and the field of chemometrics will continue to grow. This thesis is but one example, a simple harbinger of this new direction and focus. 7 .3. REFERENCES (l) Crouch, S. R. "Trends in kinetic methods of analysis" Anal. Chim. Acta 1993, 283, 453-470. (2) Crouch, S. R. "Kinetic methods of analysis - how do they rate" J. Chin. Chem. Soc. 1994, 41 , 221-229. (3) Mottola, H. A. Some Kinetic Aspects of Analytical Chemistry ; Wiley: New York, 1988. (4) Perez-Bendito, D.; Silva, M. Kinetic Methods in Analytical Chemistry ; Ellis Horwood: Chichester, 1988. (5) Perez-Bendito, D.; Silva, M. "Recent advances in kinetometrics" Trac- Trends Anal. Chem. 1996, 15, 232-240. 136 APPENDIX 137 APPENDIX MATLAB CODE FOR GENERATING SIMULATED KINETIC-SPECTROPHOTOMETRIC DATA You think you know when you learn, are more sure when you can write, even more when you can teach, but certain when you can program. --Alan J. Perlis Anyone who considers arithmetical methods of producing random digits is, of course, in a state of sin. --John von Neumann Code designed to generate simulated kinetic-spectrophotometric data was written in MATLAB’s native programming language. The main program is mulgen_a.m. This programs calls several others. Mulgen_a begins by collecting the necessary information for the generation of simulated data. The user provides a matrix of absorptivities for each reactant, reagent, and product, a matrix of the initial concentrations of each reactant, the fraction of the slower reaction to observe, the initial reagent concentration, the desired levels of instrumental noise and rate constant fluctuation, and the number of spectra to generate. The program then uses the variable perrxn (the input fraction of the slower reaction to observe), and calculates the time at which to cease “data acquisition.” 138 It does this by making the initial, worst-case assumption that the faster reaction will have reached equilibrium by the time the slower reaction has had time to begin. Using this assumption, the time at which the slower reaction will have reached the desired fractional completion is calculated. Five hundred points are then generated for the parallel reaction of the two reactants with the reagent between zero and this calculated time. The time at which the concentration of the slower product reaches the threshold determined by the fraction of the reaction to be observed is noted, and 500 more points are generated between time zero and this new time. Again, the time at which the product concentration reaches the threshold level is found, and this time is used as the stopping time for the data generation in all future calculations. The program proceeds to calculate the concentration of all species at all times in all samples. The absorptivities are then used to calculate the absorbance of each species, and these absorbances are summed to generate the final data matrix. Calculations of the concentration of the reactants, reagent, and products at any time during the reaction are performed by the program kinetic.m. This program sets up and calls the MATLAB routines for numerically solving systems of ordinary differential equations. These routines require a model of the system that includes the differential equations to be solved. This model is supplied by the function vary_k.m. The code of mulgen_a.m. kinetic.m, and vary_k.m are found below. 139 MULGEN_A function [signal,signa13,Dim_signa1,tfinal] =... mulgen_a(absorbs,concentrations,perrxn,concrO,... noisek,noisei,pts) [signal,signa13,Dim_signa1,tfinal] = mulgen_a(absorbs,concentrations,perrxn,concrO,. noisek,noisei,pts); generates absorbance data for a specified # of wavelengths over a specified time period. It uses the kinetic.m routine, and so can handle any kinetic system that can be modeled by ode45. Molar absorbtivities should be organized in columns: epsl epsZ ... epsreagent epspl epsp2 Concentrations should be also be arranged in columns. complsampl compZSampl complsamp2 comp23amp2 % % % % % % % % % % % % % % % % % % % % % % This version handles 2 components, any # of wavelengths, and % any # of samples. % It allows the user to input the amount of variation in the % rate constants, the level of instrumental noise, and the % percent completion of the slower reaction. % % It requires that k1 and k2 be declared as global variables % % The output “signal" is in the form: % % slwltl slw2t1 slw3t1 ... slw1t2 slw2t2 % % % % % % % % % % g g ‘g 1% sZwltl $2w2tl $2w3t1 ... sZwltZ sZw2t2 stltl s3w2tl s3w3tl ... s3w1t2 s3w2t2 The output "Dim_signa1" is in the form: [samples waves pts] The output I'signal3" has each row as a time, each column as a wavelength, and each sample as a page. 140 Created 9/20/95 by Tom Cullen Last Updated 6:04 PM 8/16/99 by Tom Cullen 096969690909 %*************************************************************‘k‘k % Get info from user %*************************************************************** [samples,comp]= size(concentrations); [waves,xcomp] = size(absorbs); global k1 global k2 %*************************************************************** % Determine stopping time %*************************************************************** [lowk,which_one_low]=min([k1,k2]); [highk,which_one_high]=max([k1,k2]); if kl==k2 lowk=k2; highk=k1 which_one_low=2; which_one_high=1; end medianconcs=median(concentrations); medianconc1=medianconcs(which_one_high); medianconc2=medianconcs(which_one_low); medianconcr=concr0-medianconc1-medianconc2; concratio=medianconc1./medianconc2; medianconcs=[medianconc1 medianconcZ]; reagtexcess=concr0 ./(medianconc1+medianconc2); newtf=(-log(1-(perrxn)))./ (lowk*(concrO—medianconcs(which_one_high))); [profile1,profile2,profiler] = kinetic(medianconcs,concr0,0,newtf,’vary_k’,500); if which_one_high==1 slow;profi1e=profile2; else slow;profile=profilel «end times=1inspace (0 , newtf, 500) ,- for i=1:500 141 if slow;profile(i) < (l-perrxn)*medianconcs(which_one_low) newtf=times(i); break; end; end; [profilel,profileZ,profiler] = kinetic(medianconcs,concr0,0,newtf,’vary_k’,500); if which_one_high== slow;profile=profile2; else slowgprofi1e=profilel end times=linspace(0,newtf,500); for i=1:500 if slow;profi1e(i) < (l-perrxn)*medianconcs(which_one_low) newtf=times(i); break; end; end; tfinal=newtf; %************************‘k*********'k‘k*************************** % Expand scalar data into matrices... prepare other matrices %*************************************************************** c01 = ones(samp1es,pts); c02 = ones(samp1es,pts); for i=1:pts c01(:,i) = concentrations(:,l); c02(:,i) = concentrations(:,2); end; k1_array = k1*ones(samples,pts); k2_array = k2*ones(samples,pts); k1_noisy_array = noise(k1_array,noisek); k2_noisy_array = noise(k2_array,noisek); k1_noise = abs(k1_array - k1_noisy_array); k2_noise = abs(k2_array — k2_noisy_array); noisesign = sign(rand(samp1es,pts)); time = 1inspace(0,tfina1,pts); times = ones(samp1es,pts); for i=1:samples times(i,:)=time; end; for s=1:samples for c = lzcomp eval(['con' int2str(c) ’_’ int25tr(s) ’ =... ones(waves,pts);']) eval(['conp' int23tr(c) ’_’ int2str(s) ' =... 142 ones(waves,pts);’]) eval(['abs’ int25tr(c) '_’ int23tr(s) ’ =... ones(waves,pts);’]) eval([’absps’ int25tr(c) ’_’ int25tr(s) ' =... ones(waves,pts);']) end; end; absr = ones(waves,pts); datapoints = waves*pts; signal = ones(datapoints,samples); %*************************************************************** % Calculate concentrations as a function of time %************************‘k************************************** [profilel,profile2,profiler] =... kinetic(concentrations,concr0,0,tfinal,’vary_k’,pts); k1t_noise = exp(-k1_noise .* times .* profiler) .* noisesign; k2t_noise = exp(-k2_noise .* times .* profiler) .* noisesign; profilel profilel .* k1t_noise; profile2 profileZ .* k2t_noise; profiler = profiler .* k1t_noise .* k2t_noise; profilepl = c01 - profilel; profilep2 = c02 - profi1e2; for s=1:samples for c = 1:comp for i = 1:waves eval(['con' int25tr(c) ’_’ int25tr(s) '([i],:) =... profile’ intZStr(c) ’ (s,:);']) eval(['conp’ int23tr(c) '_' int23tr(s) '([i],:) =... profilep' int25tr(c) ' (s,:);’]); eval([’conr_’ int23tr(s) ’([i],:) =... profilerls,:);’]); end; end; end; %*************************‘k'k************************************ % Convert absorptivities into a useful fonm %*************************************************************** for c = 1:comp eval([’eps’ int25tr(c) ’ = absorbs(:,’ int23tr(c) ' );']) eval(['epsp' int25tr(c) ' =... absorbs(:,' int23tr(c+comp+1) ’ );’]) end; eval([’epsr = absorbs(:,' int23tr(comp+1) ’ );’]) 143 for C: 1:comp eval([’extinct’ int25tr(c) ’ = ones(waves,pts);’]) eval([’extinctp’ int2str(c) ' = ones(waves,pts);’]) end; extinctr = ones(waves,pts); for C: 1:comp for i = lzpts eval(['extinct' int25tr(c) ' (:,[i]) =... eps’ int25tr(c) ’ (:,1);’]) eval(['extinctp' int23tr(c) ' (:,[i]) =... epsp' int23tr(c) ’ (:,1);’]) end; end; for i = lzpts extinctr(:,[i]) = epsr(:,l); end; %************************'A’************************************** % Calculate absorbances %*************************************************************** for S: 1:samp1es for C: 1:comp eval(['abs’ int23tr(c) ’ ’ int23tr(s) ' = extinct’... int25tr(c) ' .* con' int23tr(c) ' ’ int25tr(s) ';’]) eval(['absps’ int25tr(c) ’ ' int25tr(s) ' =extinctp’... int2str(c) ' .* conp’ int23tr(c) ’_' int23tr(s) ’;']) eval([’absr_’ int2str(s) ’ = conr_’ int25tr(s) ’... .* extinctr;']) end; end; for S: 1:samp1es eval([’sig_’ int23tr(s) ’ = absr_’ int23tr(s) ';’]) for C: 1:comp eval([’sig_’ int23tr(s) ’ =... sig_' int25tr(s) ’ + abs' int2str(c) ’_' int25tr(s) '... + absps' int25tr(c) '_' int25tr(s) ’ ;’]) end; end; %*************************************************************** % Reshape data and add instrumental noise %*************************************************************** for S: 1:samp1es eval(['signa1_' int23tr(s) ' =... reshape(sig_' int23tr(s) ',datapoints,1);']) eval([’signal(:,s) = signa1_’ int2str(s) ’;']) end; 144 signal noise(signal,noisei); signal = signal’; Dimhsignal=[samp1es waves pts]; signa13=reshape(signal,samples,waves,pts); signal3=permute(signa13,[3 2 1]); %**********************************‘k'k*************************** % All done!!!!! %*************************************************************** 145 KINETIC function [profilel,profilez,profiler] =... kinetic(concs,r0,t0,tfinal,model,pts) % [profilel,profi1e2,profiler] =... % kinetic(concs,r0,t0,tf,model,pts); % is a general-purpose m-file that uses the ode45 % function to generate kinetic profiles. It must be given % initial concentrations, beginning and ending times, % and the filename of the ode45 function that describes the % system. This m-file uses a cubic spline to interpolate % as it generates a user-supplied number of data points % per profile. % global k1 global k2 [mcon,ncon] = size(concs); profilel=zeros(mcon,pts); profile2=zeros(mcon,pts); profiler=zeros(mcon,pts); for 22: lzmcon c0= [concs(zz,:),r0]; [t,c]=ode45(model,[t0 tfinal],c0); ti=linspace(t0,tfinal,pts); ci=interp1(t,c,ti,’spline’); Ci=ci’; profile1(zz,:)=ci(1,:) profile2(zz,:)=ci(2,:), profiler(zz,:)=ci(3,:) I I end; 146 VARY_K function cdot=varyk(t,c); % function cp=varyk(t,c); % m-file that returns state derivatives when given state and % time valus for a second order kinetic process given by: % A+R=PA % B+R=PB % Created 8/7/96 by Tom Cullen global kl global k2 % Define cl=A, c2=B, and c3=R % cl’ = -k1 .* c1 .* c3 % c2’ = -k2 .* c2 .* c3 % c3’ = (—kl .* cl .* c3) + (-k2 .* c2 .* c3) cdot=[(-k1)*c(l)*c(3);(-k2)*c(2)*c(3);((-k1)*c(1)*c(3))+... (-k2*c(2)*c(3))]; 147 MICHIGAN S TATE UNIV. LIBRARIES lllllllllllllllllllll"Hill”llllllllllllllllll 93020582064 ll" 12