.Wa‘iJfies. : .. . .e. fiAffifiFfi . ~V A V J 355.10.. .1 . . . . 3 m . u ‘ . ‘ . . , ... 4.. . ... u. 23., 1 .. .\.é-...h.:tst. . 2.. I .1 a a a. g 2%. hr . :hzw: .. 15 .: am. you. .. fix... .5an r H.- (I 1910509. slur»... u. :33. It. r 7.3... 1... 13,311.?! $31. ., 31.3. . is: 1.115 2.2:: . . . .. 4.». .....,.u...,. fififimkh'wu - ou . v a. .r.u.‘u.u....§t.....ai it- {.3 a, 2) S. .. 12.6.1‘... 7.2.3.3. . » I»: t. .1. 1. inst: .3 All, . . x 5‘ 01).: 3: s. 2;. ...:w§.:x muses gnsm LlanAmES 1111111111111 111 11111 1111 11 1111111111 This is to certify that the dissertation entitled An Investigation of the Feasibility of Using 13 Situ ATR FTIR Spectroscopy in the Measurement of Crystallization Phenomena for Research and Development of Batch Crystallization Processes presented by Dilum D. Dunuwila has been accepted towards fulfillment of the requirements for Ph.D. degree in _Ch_emi_c_él_Engineering '_-— ,- I ‘ ' _ \1_ 7. / . K .1, . .1, 1 , 1 1 / / /' (' /.-:", , / "~':$j‘_)r I ,.; ./ . (,-< “" Major professor I l/_ Date ‘.I,"’. / / .z /1/ . lx MSU is an Affirmative Action/Equal Opportunity Institution 0— 12771 LIBRARY Michigan state University PLACE IN RETURN BOX to roman this checkout from your rooord. TO AVOID FINES return on or before date duo. DATE DUE DATE DUE DATE DUE iJC‘J 1 6 2001" I 11 1 2 0 MSU Is An Affirmative Action/Equal Opportunity Institution W m1 AN INVESTIGATION OF THE FEASIBILITY OF USING HV SI TU ATR FTIR SPECTROSCOPY IN THE MEASUREMENT OF CRYSTALLIZATION PHENOMENA FOR RESEARCH AND DEVELOPMENT OF BATCH CRYSTALLIZATION PROCESSES. By Dilum D. Dunuwila A Dissertation Submitted to Michigan State University in partial fulfillment ofthe requirements for the degree of Doctor of Philosophy Department of Chemical Engineering 1996 ABSTRACT AN INVESTIGATION OF THE FEASIBILITY OF USING IN SI TU ATR FTIR SPECTROSCOPY IN THE MEASUREMENT OF CRYSTALLIZATION PHENOMENA FOR RESEARCH AND DEVELOPMENT OF BATCH CRYSTALLIZATION PROCESSES. By Dilum D. Dunuwila Bulk Crystallization from solution is one of the most widely used unit operations in the food, pharmaceutical and chemical industries. However, a systematic approach to process development and control of crystallization processes has not been presented due to the lack of a technique capable of meastning crystallization phenomena such as solubility and supersaturation, in situ. ATR FI‘IR (Attenuated Total Reflection Fourier transform infrared) spectroscopy provides a 1mique configuration in which theinfi'aredspectrum ofaliquidphase canbeobtainedina sltn'ry without phase separation. Initially, the feasibility of the technique itself was investigated using a Micro CIRCLE® Open Boat Cell equipped with a ZnSe (zinc selenide) ATR rod. Experiments conducted with aqueous citric acid proved that ATR FTIR spectroscopy can be successfully employed to determine solubility and supersaturation. Subsequently, establishing the technical feasibility of in situ ATR FTIR spectroscopy for the measurement of crystallization phenomena was undertaken. The viability of the technique for in situ measurements was investigated using a DIPPER® 210 immersion probe manufactured by Axiom Analytical, Inc., of Irvine, California. Initial experiments conducted using aqueous maleic acid proved that ATR FTIR spectroscopy can be successfully employed to measure supersaturation, solubility and the metastable limit, in situ, with suficient accuracy and precision. Infrared spectra provide information about the chemical nature and the molecular structure of chemical systems. Changes in molecular structure are reflected in numerous ways in the IR spectrum. Consequently, the full potential of FTIR spectroscopy as a tool to both understand crystallizing systems and measm'e crystallization parameters was explored. As such the broad scape of in situ ATR FTIR spectroscopy in the field of crystallization is demonstrated. In this investigation, the versatility of in situ ATR FTIR spectroscopy is demonstrated by its applicability in research and development of batch crystallization processes and by its use for the elucidation of molecular structures in supersaturated solutions in aid of understanding crystallization phenomena. However, its potential for implementation in control of crystallization processes, although promising, remains to be proven. A methodology for the development of a control strategy is provided along with other recommendations for further investigations. iv To Savi Hydrocortisone crystallization A practice in sleep deprivation It tested his grit But, he now fears to sit Lest he sleeps through his own presentation - Peter W. Burke Pharmacia & Upjohn Company (1995) ACKNOWLEDGMENTS I would like to thank my advisor, Dr. Kris A. Berglund, for the opportunities, the freedom and the guidance he provided toward my professional growth during my tenure as a graduate student. I am grateful to my wife, Savi, for her patience and support. I am also grateful to my parents, Mallika and Robert Keerthi, and my sister, Vishani, for all the encouragement I received over the years. A special acknowledgment is due to my sister-in-law, Shirani Perera, her husband, Dr. Priyantha Perera, and their daughter, Dilshani Perera, for their hospitality and help extended to me during my early years in the United States of America. Many thanks to my colleagues, Dr. David Gagliardi, Dr. Everson Miranda, Dr. Borlan Pan, Leslie Carroll II, Sanjay Yedur, Marketta Uusi- Penttila, Chris Esswein, Klaus Weisphenig, Yilan Ling, Arvind Mathur, Laurie Ruiz, Dr. Tom Carter, Dr. Beatrice Torgerson, Mark Torgerson and Dr. Hasan Alizadeh, for making my tenure as a graduate student easier in numerous ways. I would like to thank the members of my dissertation committee, Dr. Kris A. Berglund, Dr. Alec Scranton, Dr. Robert Y. Ofoli, Dr. Stanley Crouch and Dr. Clark Radclifl'e for their time and efi'ort. A special word of thanks to Dr. Clark Radclifi‘e for his participation in this venture. I acknowledge the financial support from the United States Department of Agriculture and the Crop and Food Bioprocessing Center - Research Excellence Fund at Michigan State University. Finally, I would like to thank Dr. Mike Doyle, the President of Axiom Analytical, Inc., for his continuing support for this venture. TABLE OF CONTENTS LIST OF TABLES .................................................................................................... ix LIST OF FIGURES .................................................................................................. x CHAPTER 1 INTRODUCTION ............................................................................................... 1 1. 1 Identification and Significance of the Investigation ......................... 1 1.2 References ................................................................................................ 8 CHAPTER 2 AN INVESTIGATION OF THE APPLICABILITY OF ATR FTIR SPECTROSCOPY FOR MEASUREMENT OF SOLUBILITY AND SUPERSATURATION IN SLURRIES ....................................................... 10 2.1 Background ............................................................................................ 10 2.2 Materials and Methods ........................................................................ 13 2.2.1 Construction of the Calibration Curves .................................. 15 2.2.2 Measm'ement of Solubility Using Slurries .............................. 16 2.3 Results and Discussion ........................................................................ 17 2.4 Conclusions ............................................................................................ 30 2.5 References .............................................................................................. 3 1 CHAPTER 3 AN INVESTIGATION OF THE FEASIBILITY OF ATR FTIR SPECTROSCOPY FOR IN SI TU MEASUREMENT OF CRYSTALLIZATION PHENOMENA. ......................................................... 32 3.1 Background ............................................................................................ 32 3.2 Materials and Methods ........................................................................ 33 3.2.1 Measurement of Solubility In Situ ........................................... 35 3.2.2 Measurement of Supersaturation In Situ .............................. 35 3.2.3 Measurement of CSD ................................................................. 35 3.2.4 Construction of the Calibration Curves .................................. 36 3.3 Results and Discussion ........................................................................ 36 3.3.1 Measurement of Solubility In Situ ........................................... 38 3.3.2 Measurement of the Metastable Limit In Situ ..................... 39 3.3.3 Measurement of Supersaturation In Situ .............................. 44 3.3.4 Calibrations .................................................................................. 52 3.4 Conclusions ............................................................................................ 54 3.5 References .............................................................................................. 55 CHAPTER 4 IDENTIFICATION OF FTIR SPECTRAL FEATURES RELATED TO SOLUTION STRUCTURE FOR UTILIZATION IN MEASUREMENT OF CRYSTALLIZATION PHENOMENA ................ 56 4.1 Background ............................................................................................ 56 4.1.1 Identification of the Vibrations of the Carboxyl Group ........ 58 4.2 Materials and Methods ........................................................................ 59 4.2.1 Measurement of Solubility in Slurries ..................................... 60 4.2.2 Construction of the Calibration Curves .................................. 60 4.2.3 Data Processing ........................................................................... 60 4.3 Results and Discussion ........................................................................ 6 1 4.4 Conclusions ............................................................................................ 93 4.5 References .............................................................................................. 93 CHAPTER 5 CONTINUING INVESTIGATIONS ............................................................ 95 5. 1 Background ............................................................................................ 95 5.1.1 Control Strategy .......................................................................... 95 5.1.2 Solution Structure .................................................................... l 10 5.1.3 Effect of impurities ................................................................... 1 12 5.2 Conclusions ......................................................................................... 1 14 5.3 References ........................................................................................... 1 l4 APPENDD( ............................................................................................................ 116 LIST OF TABLES Table 2.1 Comparison of experimental solubility measurements of citric acid in water to literature values [5]. The given solubilities are based on % (w/w). ............................................. 28 Table A1 Data for Figure 2.4 ................................................................... 1 16 Table A2 Data for Figures 2.5 and 2.6 .................................................. 1 17 Table A3 Data for Figures 2.7 and 2.8 .................................................. 1 18 Table A4 Data for Figures 3.3 and 3.5 .................................................. l 19 Table A5 Data for Figure 3.6 ................................................................... 120 Table A6 Data for Figure 3.7 ................................................................... 121 Table A7 Data for Figure 3.8 ................................................................... 122 Table A8 Data for Figure 3.9 ................................................................... 123 Table A9 Data for Figure 3.10 ................................................................ 124 Table A10.1 Data for Figure 3.11 ................................................................ 125 Table A10.2 Data for Figure 3.11 ................................................................ 126 Table A10.3 Data for Figure 3.11 ................................................................ 127 Table A10.4 Data for Figure 3.11 ................................................................ 128 Table A10.5 Data for Figure 3.11 ................................................................ 129 Table A10.6 Data for Figure 3.11 ................................................................ 130 Table All Data for Figure 4.5 ................................................................... 131 Table A12 Data for Figure 4.6 ................................................................... 132 Table A13 Data for Figme 4.10 ................................................................ 133 Table A14 Data for Figm'e 4.11 ................................................................ 134 Table A15 Data for Figure 4.12 ................................................................ 135 Table A16 Data for Figure 4.13 ................................................................ 136 Table A17 Data for Figure 4.17 ................................................................ 137 Table A18 Data for Figure 4.18 ................................................................ 138 Table A19 Data for Figure 5.4 .................................................... A ............... 1 39 Figm'e 1.1 Figure 1.2 Figure2.1 Figure 2.2 Figtn'e 2.3 LIST OF FIGURES Comparison of cooling profiles for batch crystallization. These profiles are not specific to any system. They are presented to demonstrate the distinction between the three operating policies .................................................................................................. 6 Projected supersaturation profiles as a consequence of applying cooling profile given for batch crystallization. ............. 7 A schematic diagram of the operation of internal reflection spectroscopy. Infrared radiation is totally reflected at the media interface and is propagated as a transverse wave. The evanescent field generated by infrared radiation penetrates into the rarer medium in the z-direction as an exponentially decaying wave. It is composed of electric vector components in all spatial directions. IRE stands for internal reflection element [1] . ....................................................................................... 1 1 The cross section of the Micro CIRCLE® Open Boat Cell (Spectra-Tech) and its optical schematic. ATR stands for attenuated total reflection [4]. ...................................................... 14 IR spectra of citric acid in water compared to the IR spectrum of water. The arrows indicate the directions of change of the spectrum as the concentration of citric acid in water is increased. Spectra recorded here were taken at 35 °C. Spectra at both 10 °C and 30 °C also followed the same trend. .................................................................................................. 1 9 Figure 2.4 Figme 2.5 Figure 2.6 Figure 2.7 Figure 2.8 Calibration of RT,. RT1 is the transmittance ratio of the transmission band at 3277 cm'1 to that at 2610 cm". The solubilities indicated in the figure are literature values [5]. They were superimposed on the experimental calibrations to demonstrate the extension of RT! into the supersaturated region. ................................................................................................. 2 1 Calibration of RT2 at 10 °C. RT2 is the transmittance ratio of the transmission band at 3277 cm‘1 to that at 1220 cm". The solubility indicated in the figure is the literature value [5]. It was superimposed on the experimental calibration to demonstrate the extension of RT2 into the supersaturated region ................................................................................................. 22 Calibration ofRT2 at 30 °C. RT2 is the transmittance ratio of the transmission band at 327 7 cm'1 to that at 1220 cm”. The solubility indicated in the figure is the literature value [5]. It was superimposed on the experimental calibration to demonstrate the extension of RT2 into the supersaturated region. ................................................................................................. 23 Plot of RTl of slurries at equilibrium as a fimction of temperature. RT, is the transmittance ratio of the transmission band at 3277 cm‘1 to that at 2610 cm“. Each data point represents the average of measurements made using three different samples. ....................................................... 25 Plot ofRTz of slurries at equilibrium as a function of temperature. RT2 is the transmittance ratio of the transmission band at 327 7 cm'1 to that at 1220 cm". Each data point represents the average of measurements made using three different samples. ....................................................... 26 Figure 3.1 Figure 3.2 Figtn'e 3.3 Figure 3.4 Figure 3.5 Figm'e 3.6 A schematic of the experimental setup. a = spectrometer, b = ATR immersion probe, c = Crystallizer, d = Chiller, e = Crystallizer temperature relay, f = Product holding tank, g = Temperature controlled bath, h = Buchner funnel, I = vacuum ................................................... 34 ATR FTIR spectra of aqueous maleic acid as the maleic acid concentration (% w/w) was increased (a=50%, b=60%, c=65% maleic acid in water). The arrows indicate the direction of peak intensity movement with increase in maleic acid concentration. The thick arrows indicate the peaks chosen for the transmittance ratio, TR. __ transmittance at 3394 cm’1 — transmittance at 1172 cm'1 Solubility of maleic acid with respect to TR measured in Situ, in a slurry. _ transmittance at 3394 cm'1 — transmittance at 1172 cm'1 Simulated desupersaturation profiles, (a), of cooling batch crystallization. The desupersaturation profiles 1, 2 and 3 correspond to parabolic, linear and natural cooling profiles, (b), respectively ....................................................................................... 4 1 The metastable limit of maleic acid, (Y,), measured at a cooling rate of 1.3 °C/min. and a stirring rate of 420 rpm. The result of one experiment is given for illustration purposes. For the given OOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO 3i TR ................................................ 4O TR experiment, point X denotes the initial conditions and point Y1 denotes spontaneous nucleation. All points, Y,, were extracted from similar experiments. - -1 TR = transm‘ttance at 3394 cm 1 ................................................ 43 transmittance at 1172 cm' Experimental cooling profiles. Profiles 1 and 2 are parabolic cooling profiles and profile 3 is a linear cooling profile ............... 46 Figm'e 3.7 Figure 3.8 Figure 3.9 Figure 3.10 Figure3.11 Figtn‘e 4.1 Figure 4.2 Figure 4.3 The desupersaturation profile corresponding to cooling profile 1 given in Figure 3.6. transmittance at 3394 cm'1 TR = 1 1 ................................................ 47 transmittance at 1172 cm' The desupersaturation profile corresponding to cooling profile 2 given in Figure 3.6. transmittance at 3394 cm’1 1 ................................................ TR = 48 transmittance at 1172 cm' The desupersaturation profile corresponding to cooling profile 3 given in Figure 3.6. transmittance at 3394 cm'1 TR = 1 ‘1 ................................................ 49 transmittance at 117 2 cm Mass based product crystal size distributions (CSD). The CSDs 1, 2 and 3 correspond to desupersaturation profiles 1, 2 and 3, respectively. The arrows indicate the mass based average sizes of each size distribution. ....................................... 5 1 Calibrations of the transmittance ratio, TR, with respect to temperature and maleic acid concentration (% w/w maleic acid in water). The constant concentration lines were extracted from the given experimental data. transmittance at 3394 cm'1 TR = . _1 ................................................ 53 transmittance at 117 2 cm Self association of carboxylic acids via H-bonding .................... 5 7 First, second, third and fourth derivatives of a Gaussian peak ................................................................................................... 63 In situ, slurry ATR FTIR spectra of aqueous maleic acid at saturation in the low frequency region. ....................................... 64 Figure 4.4 Figure 4.5 Figure 4.6 Figure 4.7 Figm'e 4.8 Figure 4.9 Figure 4.10 First derivative profiles of the v(C=O) vibrational mode from Figure 4.3. BL1=0.005 and BL2=-0.012 were the baselines used for PIR1 and PIR2, respectively. PIR,=(DA at 1740 cm'1 - a)/(DA at 1720 cm‘1 - a) PIR2=(DA at 1740 cm’1 + b)/(DA at 1692 cm'1 + b) a=0.005 and b=0.012 are the baseline adjustments, DA=derivative absorbance ............................................................ 66 Solubility and concentration lines of maleic acid with respect to PIR, measured in situ. In addition, solubility was measured in a slurry. PIR1=(DA at 1740 cm'1 - a)/(DA at 1720 cm‘1 - a) DA=derivative absorbance a=0.005 is the baseline adjustment ............................................. 67 Solubility and concentration lines of maleic acid with respect to PH?2 measmed in situ. In addition, solubility was measured in a slurry. Pm,=(DA at 1740 cm’1 + b)/(DA at 1692 cm'1 + b) DA=derivative absorbance b=0.012 is the baseline adjustment ............................................. 69 Isolated second derivative profiles of the v(C=O) vibrational mode of aqueous maleic acid at saturation. ............................... 7 1 First derivative profiles of the 5(O-H) in-plane deformation vibrational mode from Figure 4.3. ................................................ 73 Bond distances and angles for the maleic acid molecule. Bond distances are given in A and the angles are italicized for clarity. ................................................................................................ 74 Solubility and concentration lines of maleic acid measured with respect to the peak shift of the v(C=O) vibrational mode, in situ. In addition, solubility was measured in a slurry. Absolute peak positions at two conditions are shown for reference purposes. ........................................................................................... 7 7 xiv Figure4.11 Figure 4.12 Figm'e 4.13 Figure 4.14 Figtn'e 4.15 Figure 4.16 Figure 4.17 Figure 4.18 Solubility and concentration lines of maleic acid measured with respect to the peak shift of the v(C-O) vibrational mode, in situ. In addition, solubility was measured in a slurry. Absolute peak positions at two conditions are shown for reference purposes. ........................................................................................... 7 8 Solubility and concentration lines of maleic acid measured with respect to the peak shift of the v(C=C) vibrational mode, in situ. In addition, solubility was measm'ed in a slurry. Absolute peak positions at two conditions are shown for reference purposes. ........................................................................................... 79 Solubility and concentration lines of maleic acid measured with respect to the peak shift of the 5(O-H) vibrational mode, in situ. In addition, solubility was measured in a slurry. Absolute peak positions at two conditions are shown for reference purposes. ........................................................................................... 8 1 In situ, slun'y ATR FTIR spectra of aqueous maleic acid at saturation in the high frequency region ....................................... 82 Gaussian components of the v(O—H) region. Due to the closeness of the fit the original spectral profile (normalized and baseline corrected spectrum at saturation at 30.75 °C from Figure 4.14) and the six component Gaussian function used for the fit are superimposed ................................................................. 84 Peak shift of the component near 3150 cm'1 from Figure 4.15. ....................................................................................... 86 Solubility and concentration lines of maleic acid measured with respect to the peak shift of the v(O-H) vibrational mode, in situ. In addition, solubility was measmed in a slurry. ....................... 87 Solubility of maleic acid in water with respect to % maleic acid (w/w). .................................................................................................. 91 XV Figure 5.1 Figure 5.2 Figure 5.3 Figm'e 5.4 Control policy projected to maximize the mean size of a crystal size distribution from a cooling batch crystallization process. The policy is based on maintaining a constant supersaturation defined by the IR parameters given by A(Pm) or 8(Pm). Amy?) = Pm '- Pm,sol and P112 50311;) = PIR,sol where Pm is any parameter discussed in Chapters 2, 3 and 4 in the context of measurement of supersaturation and solubility and Pm”, is the solubility with respectto Pm. ............................. 98 A simple schematic of a batch crystallizer and the primary components of the envisioned control system. T, = Inlet coolant temperature of the jacketed crystallizer, To = Outlet coolant temperature, w = Coolant flow rate, A(P,R) = Supersaturation, E = Primary controller, B = Heat removal system, D = Secondary controller, H = Heat flux from the crystallizer measured by T,, To and w. ....................................................................................... 99 Block diagrams of the cascade control system. The normalized heat removal secondary loop is illustrated in (b) .................... 10 1 A supersaturation profile of aqueous citric acid represented in terms of TIR. . -1 TIR = transmittance at 3284 cm .......................................... 103 transmittance at 1190 cm'1 xvi CHAPTER 1 INTRODUCTION 1.1 Identification and Significance of the Investigation Crystallization from solution is a widely used unit operation in the food, pharmaceutical, and chemical industries. Along with separation and purification, production of a specified crystal size distribution (CSD) is one of the primary goals of a crystallization process. The CSD affects the cost of operation of down-stream units such as filtration equipment and dryers. These are often the limiting steps in chemical manufacturing processes and significant cost reductions can be realized by creating CSDs that have favorable filtration and drying properties. Supersaturation is the driving force for both crystal nucleation and growth and, as such, controls the rate of crystallization and the resulting CSD [1, 2]. Therefore, control of crystallization processes requires in situ measurement and subsequent control of supersaturation. However, a technique that is suitable for in Situ measurements has not been fully developed. In today’s competitive environment, there is a need to implement control strategies that provide a quantitative output that can be either used by an operator or implemented in an automatic control scheme. A primary impetus for improved control comes from the pharmaceutical industry wherein nearly all products are crystallized at some point in their production. Pharmaceutical processing is usually done in a batch mode, which is often more difiicult to control than continuous processes primarily because batch processes are nonsteady state operations. Another consideration is batch-to-batch fluctuations, which can cause considerable variation in the crystallization process resulting in final product divergence. Reworking a batch that does not meet specifications incurs additional costs and opens the opportunity for contamination. Clearly, there is a need for the development of an in situ measuring device for the dynamic parameters of crystallization processes. Numerous analytical techniques for the measurement of solubility and supersaturation of solutes in liquids have been proposed [3]. They range fi'om simple residual weight determinations to radioactive tracer methods. The primary difficulty in using conventional analytical techniques (refi'actometry, interferometry, viscometry, density calibrations, etc.) is the procurement of a sample which truly reflects the dynamic process composition at process conditions. Generally, the separation of the saturated or the supersaturated solution from the crystals is required prior to analysis. Percolation, decantation and filtration are common methods used for this separation. During separation and subsequent transportation and analysis special care must be taken to maintain process conditions, primarily temperature, and to avoid adsorption of solute on any of the sampling and analytical equipment. In addition, the analytical techniques themselves pose some difficulties. For example, in refi'actometry, changes in the refractive index as a function of temperature and/or composition are small and often occur in the fourth decimal point. Interferometry, on the other hand, poses considerable procedural complexities. This investigation presents a novel method for the measurement of crystallization phenomena. The proposed technique, which is ATR FTIR (Attenuated Total Reflection Fomier transform infi'ared) spectroscopy [4] , provides a unique sampling configm'ation in which the infrared spectrum of the liquid phase of a slurry (a mixture composed of a solution saturated or supersaturated with a solute in contact with a dense suspension of the undissolved particles of the solute) can be obtained without phase separation. In ATR FTIR spectroscopy, the spectrum is characteristic of the vibrational structure of the material in intimate contact with the ATR apparatus. With regard to a slurry, the proposed approach is predicated on the assumption that effective contact can be limited to that between the liquid phase of the slurry and the ATR apparatus, thus allowing the spectroscopic investigation of the liquid phase without its separation from the suspended particles. The method has the potential to minimize dificulties encountered in sampling and analysis through in Situ measurement. Separation of the solution from the crystals is completely avoided. Initially, the feasibility of the technique itself was investigated using a Micro CIRCLE® Open Boat Cell equipped with a ZnSe (zinc selenide) ATR rod. Experiments conducted with aqueous citric acid proved that ATR FTIR spectroscopy can be successfully employed to determine solubility and supersaturation in slurries. The results are presented in Chapter 2. Subsequently, establishing the technical feasibility of in situ ATR FTIR spectroscopy for the measurement of crystallization phenomena was undertaken. The approach was aimed at exploiting recent developments in ATR FTIR spectroscopy for in situ measurement of supersaturation for the purposes of analysis and control of crystallization processes. The results of in situ measurement of supersaturation, solubility and the metastable limit of aqueous maleic acid are presented in Chapter 3. In addition, a simple way to extract the transient slurry density is outlined. The ATR apparatus used in this study was the DIPPER® 210 immersion probe manufactured by Axiom Analytical, Inc. of Irvine California, USA Infrared spectra provide information about the chemical nature and the molecular structure of chemical systems [5]. Infrared active molecular vibrations absorb infiared radiation, thus generating a characteristic IR spectrum. Reactions, changes in composition, introduction of new species and temperature affect the molecular vibrational structure of a chemical system. These changes are reflected in numerous ways in the IR spectrum, the most prominent being peak intensity changes generally associated with composition [6, 7]. Changes in molecular vibrational structure are manifested in more subtle spectral features such as shifts in peak positions [8]. These features, in addition to providing valuable insight to the organization of solute and solvent molecules in supersaturated solutions, can be extremely useful as a tool for measuring crystallization parameters. This is particularly true with systems that have overlapping IR bands where intensity changes are dificult to isolate. The discussion in Chapter 4 focuses on methods available to identify and isolate subtle effects of IR spectra toward elucidation of solution structure in supersaturated solutions and toward utilization in measurement of crystallization parameters. Both derivative spectroscopy [9] and deconvolution [9] of overlapping bands were used to isolate IR absorption bands arising fi'om various vibrational modes of maleic acid sensitive to reorganizations in solution structure. Several parameters suitable for measurement of solubility and supersaturation were identified and the results are presented. In addition, some insight to the organization of aqueous maleic acid in supersaturated solutions is provided. Batch crystallizers operated under natural cooling are known to produce a supersaturation peak at the onset of nucleation leading to high rates of nucleation and excessive fines formation [1, 10]. This phenomenon leads to fouling problems, reduced product yields and problems in downstream product handling and processing that can be extremely detrimental to the productivity of the process. The conventional method for circumventing this phenomenon is to use programmed cooling methods [1, 10, 11], as illustrated in Figure 1.1, where the crystallizing system responds to a programmed cooling profile. The predicted time course of the corresponding supersaturation profiles are given in Figure 1.2 [10]. However, this is an open-loop control method, in that, it does not allow intervention to optimize the output through control of supersaturation, which is the driving force for crystal nucleation and growth In large scale industrial processes, batch-to-batch variations notwithstanding, supersaturation can undergo random fluctuations due to minor changes in operating conditions. What is desired is a closed-loop control scheme that responds to fluctuations in supersaturation by adjusting the rate of cooling and/or the rate of evaporation to maintain optimal conditions during the operation. ATR FTIR spectroscopy provides an opportunity to investigate Temperature (°C) Figure 1.1 I I iii I 60 +Parabolic 55 .’. .1 . +Linear ' +Natural 50 {- 45 4. 40 {- 35 {- 30 i 25 {- 0 36 72 108 144 180 Time (min) Comparison of cooling profiles for batch crystallization. These profiles are not specific to any system. They are presented to demonstrate the distinction between the three operating policies. l 1 I j ' U I I ' ' ' +Parabolic cooling ‘ ‘11— Linear cooling 0,012 ~- +Natural cooling .4- 0.016'v-:-.. 0.008 Supersaturation 0.004 ' Figure 1.2 Projected supersaturation profiles as a consequence of applying cooling profile given for batch crystallization. the possibility of using in situ measurement of supersaturation in closed- loop control schemes for industrial crystallizers. An outline for a possible control strategy is presented in Chapter 5. At this time development of a control strategy for batch crystallization is recommended. The impetus for the focus on batch crystallization is derived from the following considerations; the unsteady state operation of batch crystallizers requires more robust control, current trends in the competitive specialty chemical industry call for the development of batch crystallization and the lack of attention batch crystallization has received in general over time. The scope of in situ ATR FTIR spectroscopy in crystallization is not limited to its use in measurement of crystallization parameters. IR spectra can provide valuable information to deduce the solution structure of supersaturated solutions of various systems. A discussion on this topic is given in Chapter 4 and is continued in Chapter 5. In addition, its potential to study effects of impurities on various systems is discussed. 1 .3 References 1. Mullin, J. W., “Crystallization”, Butterworth-Heinemann, Oxford, England, (1993). 2. Randolph, A D.; Larson, M. A, “Theory of Particulate Processes”, Academic Press, New York, USA., (1988). 3. Zimmerman, H. K, Chemical Reviews, 51 (1952) 25. 4. Mirabella, Jr., F. M., Ed., “Internal Reflection Spectroscopy”, Marcel Dekker, Inc., New York, USA, (1993). 5. Ingle Jr., J. D., Crouch, S. R., “Spectrochemical Analysis”, Prentice Hall, Englewood Cliffs, New Jersey, (1987). 10. 11. Dunuwila, D. D., Carroll II, L. B., Berglund, K A, J. Cryst. Growth, 137, (1994), 561. Dunuwila, D. D., Berglund, K. A, Revised for publication in Ind. Eng. Chem. Fundamentals, (1995). Avram, M., Mateescu, GH. D., “Infrared SpectroscopyApplications in Organic Chemistry”, Wiley-Interscience, New York, New York, (1966), Ch.7. Perkin Elmer IR Data Manager, User’s Manual, (1991). Jones, A G., Mullin, J. W., Chem. Eng. Sci, 29, (1974), 105. Mullin, J. W., Nyvlt, J ., Chem. Eng. Sci, 26, (1971), 369. * CHAPTERZ AN INVESTIGATION OF THE APPLICABILITY OF ATR FTIR SPECTROSCOPY FOR MEASUREMENT OF SOLUBILITY AND SUPERSATURATION IN SLURRIES 2.1 Background Attenuated total reflection spectroscopy is based on the presence of an evanescent field in an optically rarer medium (lower refi'active index) in contact with an optically denser medium (higher refractive index) within which radiation is propagated due to total internal reflection [1]. Figure 2.1 is a schematic of this phenomenon. The placement of an absorbing optically rarer medium in contact with the denser propagating medium facilitates the interaction between the evanescent field and the absorbing medium. The evanescent wave is a component of the propagating electromagnetic radiation. Electromagnetic radiation conveys information characteristic of matter interacting with its components. This is the basis of most spectroscopic techniques regardless of the mode of operation (transmission, internal reflection, etc.). Consequently, information characteristic of the absorbing medium is conveyed by the propagating radiation. For the current application, propagating infrared radiation generates the characteristic infrared spectrum of the absorbing medium. ’ Dunuwila, D. D., Carroll 11, 1.3., Berglund, K. A, J. Cryst. Growth, 137, (1994), 561. 10 11 a" I ..... .- o'.’ o.- 9" n.- ; Reflected IRE .; ------- 5 Wave r: .. """ Z Rarer Medium 1 Evanescent (Sample) Wave .- .o ’0 O .0 .0 o I a... .o I .o .- .- I .- D .- I. I.- ..... .. a" I O n... Figure 2.1 A schematic diagram of the operation of internal reflection spectroscopy. Infrared radiation is totally reflected at the media interface and is propagated as a transverse wave. The evanescent field generated by infrared radiation penetrates into the rarer medium in the z-direction as an exponentially decaying wave. It is composed of electric vector components in all spatial directions. IRE stands for internal reflection element [1]. 12 The evanescent field decays exponentially in the z-direction and therefore its effect is confined to the immediate vicinity of the media interface (2-5 um). It follows that intimate contact between the medium subjected to analysis and the propagating medium is essential for reliable spectral analysis. The penetration of the exponentially decaying energy field is assessed by a parameter called the depth of penetration (dp) [2]. The depth of penetration is related to the angle of incidence (0) of the propagating radiation on the internal reflection element (IRE), the wavelength (A) and the refractive index of the analyte medium relative to the denser propagating medium (am) as follows; (1,, decreases as Oincreases dp decreases as 2. decreases dp decreases as n1,1 decreases (n2,< 1) The extension of this technology for the measurement of crystallization phenomena in slurries was based on the expectation that there will be minimal contact between the crystals of a slurry and the IRE. Minimal contact between the crystals of a slurry and the IRE minimizes the interaction between the crystals and the evanescent field allowing the accumulation of solution phase properties without phase separation with little or no interference from the crystals. It could be postulated that the interaction of the evanescent field is limited to the solution phase that wets the IRE which in turn serves as a minute solution phase barrier that hinders close contact between the crystals and the IRE. In this scenario, 13 the thickness of the wetting solution phase film is expected to be greater than the depth of penetration of the evanescent field. In addition, d}D can be decreased by manipulating 9, land n21. For example, an alternate ATR configuration that has a more favorable angle of incidence can be used. Also spectral analysis can be conducted at lower wavelengths and/or an IRE with a higher refractive index can be utilized. 2.2 Materials and Methods The applicability of ATR technology toward determination of solubility and supersaturation of solutes in solution using slurries was investigated using a Micro CIRCLE® Open Boat Cell equipped with a ZnSe ATR rod manufactured by Spectra Tech Inc, Stamford, CT, USA. A schematic of the CIRCLE® cell is given in Figure 2.2. This configuration was particularly suited for preliminary experiments due to its open boat construction into which solutions or slurries can easily be dispensed. Temperature control was accomplished using a stainless steel heating jacket designed to specifications that provide best contact between the CIRCLE® cell and the jacket. The heating jacket was incorporated into the structure of the Perkin-Elmer base plate designed specifically for the use of the CIRCLE® cell. The jacket was constructed at the machine shop of the Division of Engineering Research, Michigan State University. The heating jacket was connected to a Lauda Brinkmann Refrigerating Circulator R06. The infrared spectrometer was a Perkin-Elmer 1750 14 / EE‘.--- \ A $ ‘ \;---) Cone T Micro Circle Open Boat Cell Figure 2.2 The cross section of the Micro CIRCLE® Open Boat Cell (Spectra-Tech) and its optical schematic. ATR stands for attenuated total reflection. [4]. 15 Infrared Fourier Transform Spectrometer connected to a Perkin-Elmer 7700 Professional Computer. The solute-solvent system of choice was citric acid in water. Crystalline citric acid monohydrate was obtained from Columbus Chemical Industries Inc., Columbus, WI, USA. 2.2.1 Construction of the Calibration Curves Appropriate amounts (weight) of citric acid monohydrate and distilled water were placed in 20 ml Kimble disposable scintillation vials. The mixtures were gently heated with a Master Heat Gun and stirred with a Vortex Genie 2 Mixer, alternately, to homogeneous solutions. The CIRCLE® cell was heated to the set temperature (10 °C and 30 °C) in the heating jacket. The cell and the light transfer optics were aligned to obtain an optimum energy throughput. The spectrum due to the empty cell was scanned and stored as the background. Consequently, the sample spectrum appears free of the background spectral information. The solutions were dispensed into the cell using Pasteur pipettes. Solutions were left standing in the cell for about 10 minutes to allow the solution temperature to equilibrate with the set temperature (jacket temperature). It was noted however, that less than 5 minutes was required for the temperatures to come in to equilibrium. It was also noted that, at thermal equilibrium, for all practical purposes, there was no difference in temperature between the sample contained in the cell and the heating jacket. The temperature within the cell was measured using an Omega OL-703 linear response thermistor probe connected to an Omega Digicator digital readout. 16 Sample spectra were recorded at the set temperature. The percent transmittance at 327 7 cm“, at 2610 cm'1 and at 1220 cm‘1 were recorded over a wide range of citric acid concentrations in water (approximately 0 - 70 % (w/w)). Data were collected at set temperatures 10 °C and 30 °C. Between analysis of different samples the cell was rinsed first with distilled water and then with HPLC grade acetone (obtained from Mallinckrodt Specialty Chemical 00., Paris, KY, USA). The acetone rinse was to facilitate faster drying of the cell in preparation for the next sample analysis. 2.2.2 Measurement of Solubility Using Slurries A 150 ml batch reactor equipped with a heating jacket connected to the Lauda Brinkmann circulator (same as that connected to CIRC LE® cell heating jacket) was used to prepare slurries of citric acid and water. The slurries were stirred with a Curtin Matheson Scientific 244-793 magnetic stirrer. Slurries at solubility were prepared by dissolving an excess amount of citric acid (an amount exceeding the anticipated solubility at a given temperature) in distilled water, heating the mixture to a homogeneous solution, cooling the solution to the set temperature to facilitate nucleation and stirring for 24 hours to provide sufficient time for growth and equilibration. The temperature settings ranged fium 10 °C to 35 °C. The preparation of the FTIR spectrometer and the CIRCLE® cell was identical to the procedure described under the preceding subtitle. The slurries at equilibrium were transported to the cell with a syringe (a 10 ml disposable syringe was cut at the neck to provide a wider aperture that l7 facilitated easy uptake and disposal of the slurry). Three slurry samples were analyzed at each temperature. 2.3 Results and Discussion According to the absorption law [3] the absorptance, a, can be written as; a = (1—T) = 1—e""’ (2.1) where T is the transmittance, k is the absorption coeficient and b is the thickness of the absorbing medium. Note that Equation 2.1 is valid for both absorption at an interface or surface and absorption within a medium. Equation 2.1 can be solved to give; T = e’“ (2.2) In ATR spectroscopy the equivalent expression for the path length, b, for absorption across an IRE interface is given by the product [4]; b = Nd p (2.3) where N is the number of reflection points on the surface in contact with the sample and d, is the depth of penetration. A parameter that is often used for quantitative measurements when using spectroscopic techniques for analysis is a ratio between peaks that reflect the relative changes of the system. Peak-ratioing provides an internal standard that can effectively eliminate errors due to imprecise alignment of optical coupling attachments in analyzing subsequent samples. Errors due to instrumental drifts are also minimized. There is evidence to suggest that when peak-ratioing techniques are employed for l8 analysis, calibrations developed using a particular IRE and a particular angle of incidence of the radiation on the IRE, are equally valid for an alternate IRE and an angle of incidence [1]. Let Tl be the transmittance of a characteristic infrared transmission band of a solute dissolved in a given solvent. Let T2 be the transmittance of a characteristic infrared transmission band of the solvent. Equations 2.2 and 2.3 yield; T2 _ Ndpikl-kz) — — 8 T1 (2.4) where Itl and 122 are the absorption coefficients of the solute and solvent bands respectively. The dependence of d p on wavelength has been neglected for the purpose of demonstration. The relative transmittance, RT = Tle,, given by Equation 2.4, predicts an exponential gain in RT as the concentration of the solute is increased (k1 is proportional to the solute concentration). Therefore, RT is particularly suitable for measurement of crystallization parameters since linear changes in concentration are expressed by exponential gains in RT providing substantial signal gain for the measurement. In addition, a measurement with minimum error is provided by the peak-ratio. In order to determine the solubility of citric acid in water it was necessary to develop a calibration curve of RT vs. concentration of citric acid in water. The calibration parameters were selected by observing the relative spectral dynamics of the solute/solvent system. FTIR spectra of aqueous citric acid solutions are given in Figure 2.3. The decrease in intensity of infiared transmission bands of citric acid 19 100'....5....!....5....£-....L.... o % (w/w) Citric Acid 3 80 i I. Q) Q . “ 1 c6 4.) ECO-1H ii- m b d S: 1 . CU . 1 . @401 1 .- 50 D d 1 70% 1 j 20-. ] '._. ‘4 1 o ------------------- u ~ , l l l r A i n ‘ i ‘ 3700 3200 2700 2200 1700 1200 700 Wavenumber (cm‘l) Figure 2.3 IR spectra of citric acid in water compared to the IR spectrum of water. The arrows indicate the directions of change of the spectrum as the concentration of citric acid in water is increased. Spectra recorded here were taken at 35 °C. Spectra at both 10 °C and 30 °C also followed the same trend. 20 relative to that of water (in the vicinity of 3300 cm": OH stretching) is clearly evident. Spectra at 10 °C and at 30 °C resemble those at 35 °C given in Figure 2.3. Using the observed infrared spectra two calibration parameters, RT, and RT,, were defined; transmittance of the water band at 3277 cm'1 transmittance of the citric acid band at 2610 cm’ RT _ transmittance of the water band at 3277 cm‘1 2 transmittance of the citric acid band at 1220 cm’1 RTl = 1 (2.5) As previously discussed, the depth of penetration of the evanescent field into the sample is proportional to the wavelength of the radiation. As a result, the possibility exists for the deeper penetrating energy fields at higher wavelengths to interact with the undissolved particles. Therefore, analysis at higher wavelengths could produce erratic data due to spectral contributions from the solid state. To test the consequences of the wavelength dependence of dp, the calibration parameters given in Equation 2.5 were defined; one at a shorter wavelength, RT,, and the other at a longer wavelength, R T2. The plots of RT, vs. concentration of citric acid in water are given in Figure 2.4. The temperature dependency of RT is expected due to intermolecular and intramolecular interactions such as hydrogen-bonding. The filled circles in Figure 2.4 represent literature solubilities of citric acid in water [5] and they were incorporated into the figure to illustrate the extension of the calibration curves into the supersaturated region. Figure 2.4 confirms the exponential gain in RT predicted by Equation 2.4. Plots of RT2 vs. concentration of citric acid in water are given in Figures 2.5 and 2.6. The two calibrations, at 10 °C and at 30 °C, are not 21 0.6 .fi+..flt....:m14-11”“, e +10°C ti 051: +30°C 1. “_ ' . C Solubility 11' . J I 1 n '4 1 . '{I 1 0.4.: '1. .1. . :loo . E: 0.3-1 .- .‘. m i- ..r‘ '0 i J :- :'.’. ‘ 0.211 . .. i. F . ' l D .p. . .o 1 0.1-1: "0; ._ " 'IOO. . r: :. '0 1 0 was 0 10 20 30 40 50 60 70 % (w/w) Citric Acid in Water Figure 2.4 Calibration of RT,. RT, is the transmittance ratio of the transmission band at 3277 cm‘1 to that at 2610 cm". The solubilities indicated in the figure are literature values [5] . They were superimposed on the experimental calibrations to demonstrate the extension of RT, into the supersaturated region. 22 L I l l L l ftuv'vuuv—IWYVY'YTVII'VVIII-fiv'v 0 10 °C C Solubility -.-.l.--- RT 0 10 20 30 40 50 60 70 % (w/w) Citric Acid in Water Figure 2.5 Calibration of RT, at 10 °C. RT, is the transmittance ratio of the transmission band at 3277 cm‘1 to that at 1220 cm". The solubility indicated in the figure is the literature value [5]. It was superimposed on the experimental calibration to demonstrate the extension of RT, into the supersaturated region. 23 _ —e—30°C I °f 6 -_- 1 C Solubility RT % (w/w) Citric Acid in Water Figure 2.6 Calibration of RT, at 30 °C. RT, is the transmittance ratio of the transmission band at 3277 cm'1 to that at 1220 cm". The solubility indicated in the figure is the literature value [5]. It was superimposed on the experimental calibration to demonstrate the extension of RT, into the supersaturated region. 24 presented in the same figure due to considerable overlap of data. Here too, the data confirms the predicted exponential behavior of RT. The reproducibility and the consistency of both calibration parameters near and above the solubility of citric acid in water makes them viable candidates for the measurement of solubility and supersaturation. In addition, the exponential gain in RT provides more gain for the measurement of supersaturation than concentration-based approaches. The traditional solubility and supersaturation measurement techniques (refractometry, interferometry, viscometry, density calibrations, etc.) measure a single calibrated parameter. However, IR spectroscopy, which reflects the vibrational structure of the analyte, can provide more than one piece of information. For example, in addition to solubility or supersaturation, the presence of additives or impurities can be monitored. Reaction dynamics of reactive crystallization systems can be monitored. The advantage is that necessary information can be obtained simultaneously using the same technique without resorting to multiple techniques that are variable specific. The broader scope of IR spectroscopy in crystallization is taken up in the chapters to follow. RT, and R T, of slurries at equilibrium were plotted against temperature in Figures 2.7 and 2.8, respectively. They are the solubility curves of aqueous citric acid with respect to RTs. The solubility of aqueous citric acid with respect to % MW at 10 °C and at 30 °C was determined using RT, and RT, of slurries at equilibrium and the respective calibration 25 -1- 0.55 n W, 1 0.5 0.45 sic-4 0.35 0.3 0.25;----;+1--,--U,UH:,_H: 1-- 1 5 10 15 20 25 30 35 40 Temperature (°C) Figure 2.7 Plot of RT, of slurries at equilibrium as a function of temperature. RT, is the transmittance ratio of the transmission band at 3277 cm" to that at 2610 cm“. Each data point represents the average of measurements made using three different samples. 26 11 1 . ;. iw :1 1‘ 10 J- 1. Q'E :11- 81:. .. e" -- 1. a: 7 1 1 6.: .. 4:-b 1. 3’ A: u: U: A : A: -1 ‘ 5 10 15 20 25 30 35 40 Temperature (°C) Figure 2.8 Plot of RT, of slurries at equilibrium as a function of temperature. RT, is the transmittance ratio of the transmission band at 3277 cm“1 to that at 1220 cm". Each data point represents the average of measurements made using three different samples. 27 curves at the respective temperatures. Solubilities thus determined are compared to literature values [5] in Table 2.1. The percent deviation of the solubility determined using RT, from the literature value is relatively small. These deviations are most likely due to inadequate control in experimental procedure rather than due to the analytical technique. Since this investigation, at this stage, was primarily concerned with proving the viability of the concept, precise control of the experiment was not implemented. For example, disturbances due to temperature fluctuations and solute adsorption on the sampling apparatus during sampling of slurries were not controlled. A difference of 1 °C between the reactor temperature and the CIRCLE® cell temperature was observed. For these experiments the digital temperature readout of the Lauda Brinkmann circulator and the digital readout connected to the thermistor probe were not calibrated against a standard thermometer. Such procedural inadequacies were most likely responsible for the minor deviations in measurement. The measurement using RT, overestimates solubility by 3.7 % at 10 °C and by 10 % at 30 °C. It is apparent fi‘om Figure 2.8 that the solubility measured using RT, increasingly overestimates the solubility at higher temperatures. As discussed before, RT, is a parameter that probes deeper into the slurry compared to RT,. It is possible that significant contributions fi-om the crystals of the citric acid slurry to the solution state spectrum could result in an overestimation. However, it is not clear why overestimation is increasingly higher at higher temperatures. Scattering 28 Table 2.1 Comparison of experimental solubility measurements of citric acid in water to literature values. [5]. The given solubilities are based on % (w/w). Solubility - Solubility Measurement “1115,65?th ExSp ($133,116,331 Given in the % Deviation Parameter Literature RT, 10 55.8 54.0 + 3.3 RT, 30 63.6 64.3 - 1.1 RT, 10 56.0 54.0 + 3.7 RTz 30 70.7 64.3 + 10.0 29 of data is also increasingly higher at higher temperatures. A likely cause is not clear at this moment. If the deeper penetration of the evanescent field into the sample was in some way responsible for the overestimation of solubility and scattering of data, the employment of a greater angle of incidence for the radiation on the IRE and an IRE with a higher refractive index (Ge and AMTIR-l ATR elements) could help overcome such problems through the reduction of the depth of penetration. Such parametric manipulations may be absolutely necessary for solute-solvent systems that have transmission bands of utility only at higher wavelengths. Concentrations of supersaturated solutions of citric acid in slurries were not measured using the CIRCLE® cell. The degree of control in the crude sampling procedure followed in these experiments were deemed inadequate to study the labile supersaturated region. It was a limitation imposed by the procedure and certainly not by the technology. However, in recognition of the feasibility of the technique for the measurement of supersaturation, a few comments follows. Supersaturation, S, is defined by the following equation [6]; S=nln{ YC } (2.6) quCeq where 7 is the activity coefficient, C is the concentration and n is the number of ions in a molecular unit. The subscript "eq" denotes the solute- solution equilibrium. The activity coefficient, 7, reflects nonideality of solutions. Generally, nonideal behavior is due to extensive intermolecular 30 and intramolecular interactions between the solute and the solvent. These interactions in turn configure the solution structure of the system. Then, it follows that the activity coefficient is reflective of the solution structure at a molecular level. Therefore, a method for determining supersaturation that is reflective of the solution structure would be inherently more accurate. However, due to practical restraints in measuring activities, conventional techniques such as refractometry, interferometry, viscometry and densitometry that measure the respective bulk properties are used along with simplifying assumptions. However, IR spectroscopy, which probes the vibrational and rotational structure of molecules is reflective of the solution structure and therefore is a more viable technique to measure supersaturation that is more reflective of activities. Recently, other solution structure based methods such as fluorescence spectroscopy have been used to measure supersaturation in aqueous systems [7, 8]. 2.4 Conclusions This investigation clearly demonstrates the feasibility of ATR FTIR spectroscopy toward measurement of solubility and supersaturation in slurries. The exponential gain of the defined parameters, RT, and RT,, with increasing solute concentrations provides increased signal gain for more accurate data analysis. The capability to carry out measurements using slurries certainly eliminates the separation of phases required by other measurement techniques. The feasibility of extending ATR FTIR spectroscopy for in situ measurement of crystallization phenomena such as solubility and supersaturation is explored in the following chapter. The 31 noted inconsistency in solubility measured using RT, and RT, is addressed in Chapter 4. 2.5 References 1. Mirabella, Jr., F. M., in: “Internal Reflection Spectroscopy”, Ed. F. M. Mirabella, Jr., Marcel Dekker, Inc., New York, USA, (1993), ch. 2. Harrick, N. J ., duPre, F. K., Appl. Opt, 5, (1966), 1739. Ingle, Jr., J. D., Crouch, S. R., “Spectrochemical Analysis”, Prentice Hall, Englewood Cliffs, USA, (1988), ch. 3. Coates, J. P., in: “Internal Reflection Spectroscopy”, Ed. F. M. Mirabella, Jr., Marcel Dekker, Inc., New York, USA, (1993), ch. 3. The Merck Index, Ed. S. Budavari, Merck and Co., Inc., Rahway, USA, (1989), 11th edition. Garside, J ., Chem. Engng. Sci, 40:1, (1985), 3. Chakraborty, R., Berglund, K. A, AIChE Symp. Ser., 88:284, (1991), 114. Chakraborty, R., Berglund, K A, J. Cryst. Growth, 125, (1992), 81. * CHAPTER3 AN INVESTIGATION OF THE FEASIBILITY OF ATR FTIR SPECTROSCOPY FOR IN SI TU MEASUREMENT OF CRYSTALLIZATION PHENOMENA 3.1 Background Numerous analytical techniques for the measurement of supersaturation of solutes in liquids have been proposed [1, 2, 3, 4, 5, 6]. They range from simple residual weight determinations to radioactive tracer methods. The primary difficulty in using most conventional analytical techniques for the measurement of supersaturation is the requirement of phase separation for analysis. Both the presence of a slurry and the shiny density can adversely affect the measurement. Consequently, such techniques are not adequately suited for in situ measurement. ATR FTIR spectroscopy [7] provides a unique sampling configuration in which the infi'ared spectrum of the liquid phase of a slurry can be obtained without phase separation. The approach was predicated on the assumption that effective contact can be limited to that between the liquid phase of the slurry and the ATR apparatus. The validity of the assumption was verified in the study presented in Chapter 1 [8]. Substantial evidence for the feasibility of ATR FTIR spectroscopy for the measurement of ' Dunuwila, D. D., Berglund, K. A, (in review) Ind. Eng. Chem. Fundamentals, (1995). 32 33 solubility and supersaturation was provided. However, the Micro CIRCLE® Open Boat Cell used in the study presented in Chapter 2 was not configured for in situ measurements. Therefore, the focus of this chapter is establishing the technical feasibility of ATR FTIR spectroscopy for in situ measurement of crystallization phenomena. This continuing effort was inspired by recent developments in ATR FTIR technology in the form of various flexible radiation transfer systems and ATR configurations particularly suited for in situ measurements. One such device is the DIPPER® 210 immersion probe manufactured by Axiom Analytical, Inc. of Irvine California, USA that was used in this investigation. 3.2 Materials and Methods In situ measurements of supersaturation, solubility and the metastable limit were performed using a DIPPER® 210 ATR FTIR immersion probe equipped with an AMTIR- 1 conical internal reflection element manufactured by Axiom Analytical, Inc. of Irvine, California, USA. A simple schematic of the experimental setup is given in Figure 3.1. Temperature control of the 2L baflled and jacketed glass crystallizer was accomplished by a Lauda Brinkmann Refrigerating Circulator, model RC6. The crystallizer was pln'chased from Lab Glass, Inc. Stirring was provided by a 2" polystyrene marine impeller powered by a Fisher Scientific StedFastn‘ stirrer, model SL600. The infrared spectrometer was a Perkin- Elmer 1750 Infrared Fourier transform spectrometer connected to a 34 Figtn‘e 3. 1 A schematic of the experimental setup. a = spectrometer, b = ATR immersion probe, c = Crystallizer, d = Chiller, e = Crystallizer temperature relay, f = Product holding tank, g = Temperature controlled bath, h = Buchner funnel, I = vacuum 35 Gateway 2000 486DX2-50V micro computer and the spectrometer was supported by Perkin-Elmer IRDM software. The slmry temperature was recorded using an Omega RD-TEMP-XT temperature logger and monitored umng an Omega OL-7 03 linear response thermistor probe connected to a Omega DIGICAT0R® digital readout. The solute-solvent system studied was maleic acid in water and maleic acid was purchased from the Aldrich Chemical Company. 3.2.1 Measurement of Solubility In Situ Solubility was measured in a slurry at temperatures ranging fi'om 50°C to 67°C. At each temperature, solubility was approached both fi'om supersaturation and undersaturation in order to insure accuracy of the measm'ements. At least five spectra were scanned per setting (i.e., at least 10 spectra at each temperature setting since solubility was approached both from supersaturation and undersaturation). 3.2.2 Measurement of Supersaturation In Situ A 63% (w/w) maleic acid solution (total weight was 2 kg) was slightly supersatm'ated by cooling to 63°C, seeded and cooled by applying predetermined cooling profiles. Two grams of 463 micron seeds were used. Three distinct cooling profiles were used. The product crystals were collected at 56°C. ATR FTIR spectra were accumulated every 20 seconds. 8.2.3 Measurement of CSD The product crystals were vacuum filtered and washed with ethyl ether under vacuum immediately. The product was dried overnight at 45°C in a Fisher Scientific ISOTEMP programmable oven, model 818F, before subjecting to sieve analysis. Sieve analysis was done using a Tyler RX-86 36 sieve shaker. 3.2.4 Construction of the Calibration Curves Appropriate amounts (weight) of maleic acid and distilled water were placed in the crystallizer and heated to a homogeneous solution. The solution was cooled while accumulating solution ATR FTIR spectra, in situ, every 2 minutes until maleic acid spontaneously nucleated. A series of such experiments ranging from 59 to 65 % (w/w) maleic acid was conducted. At each weight percent, experiments were conducted in duplicate. 3.3 Results and Discussion ATR FTIR spectra of aqueous maleic acid solutions are given in Figm'e 3.2. The decrease in intensity of infiared transmission bands of maleic acid relative to that of water (in the vicinity of 3400 cm*1: OH stretching) as the maleic acid concentration is increased fi‘om 50 to 65 %(w/w) is clearly evident. In Chapter 2, the merits of using a transmittance intensity ratio between a solvent band and a solute band as the entity that reflects crystallization parameters were discussed and results that justify the use of a transmittance ratio were presented [8]. Accordingly, we used, a transmittance intensity ratio for in situ measurement of supersatm'ation, solubility and the metastable limit. Technically, it is plausible to use any of the maleic acid bands that appear in Figure 3.2 as the denominator of the transmittance ratio. However, upon closer observation it was noted that the band at 1 172 cm'1 is more sensitive to changes in solute concentration than are the bands at 1222 cm'1 and 1707 ch. Therefore, the 37 80‘ 70 60 50 % Transmittance 40‘ 3800 3400 3000 2600 1700 1500 1300 1100 Wavenumber (cm-1) Figure 3.2 ATR FTIR spectra of aqueous maleic acid as the maleic acid concentration (% w/w) was increased (a=50%, b=60%, c=65% maleic acid in water). The arrows indicate the direction of peak intensity movement with increase in maleic acid concentration. The thick arrows indicate the peaks chosen for the transmittance ratio, TR. _ transmittance at 3394 cm'1 TR " . -1 transmittance at 1172 cm 38 transmittance ratio, TR, was defined as; _ transmittance at 3394 cm'1 _ (3.1) transmittance at 117 2 cm'1 TR In the following discussion, we have chosen to represent solubility, supersaturation and the metastable limit in terms of TR since TR, which is a function of concentration and temperature, reflects these parameters implicitly. Alternatively, crystallization parameters can be represented in terms of concentration. In doing so, however, the useful signal gain provided by the exponential gain in T, is sacrificed. 3.3.1 Measurement of Solubility In Situ Achievement of equilibrium poses a significant problem in measurement of solubility. Often, prolonged agitation at constant temperatme is necessary for systems to reach equilibrium. Contact for days and weeks is not uncommon for some highly viscous solutions and systems at relatively low temperatures. Substances of low solubility may also require long contact times. Thus, the most convenient method for measurement of solubility would be one that is suitable for in situ measurements since a system approaching equilibrium can be monitored, minimizing expended time and effort. As alluded to in the previous chapter, conventional techniques do not meet these requirements. For example, it is inconvenient to monitor a system steadily approaching equilibrium using these techniques since portions of solutions have to be sampled fiequently for analysis. However, continuous observation can be easily accomplished through in situ utilization of ATR FTIR spectroscopy. In order to check the accuracy of solubility determinations, it is recommended to approach 39 equilibrium from both the undersaturated and supersaturated states. The capability of an in situ measuring device to continuously monitor the system dynamics make such checks easier. The solubility of maleic acid in water measured in a shiny, in situ, is given in Figure 3.3. The data represent the average of T, obtained upon approaching solubility from both the supersaturated state and the undersaturated state. In both approaches, when there was no significant change in T,, the system was considered to be at or close to equilibrium. Since the measured solubility can be the average of measurements fi'om both approaches, it is not necessary to prolong experiments lmtil thermodynamic eqm'librium is reached. Thus, the technique is convenient and time-saving. The experimental solubility data clearly demonstrate that ATR FTIR spectroscopy can be used to measure crystallization parameters in situ, in slurries, with a tolerable noise level. Certainly, these features of in situ ATR FTIR spectroscopy manifests the viability of the technique for application in process development of crystallization processes. 3.3.2 Measurement of the Metastable Limit In Situ The metastable limit along with the solubility defines the operating zone of bulk crystallization fi'om solution as illustrated in Figure 3.4. The metastable limit is the critical supersaturation at which point the system nucleates spontaneously (primary nucleation) [9]. Therefore, the dominating event of a system operating in the vicinity of the metastable limit is primary nucleation. Primary nucleation generates an excessive 40 4.5 . . . - AT 52 54 56 58 60 Temperature (°C) I U 62 64 66 68 Figm'e 3.3 Solubility of maleic acid with respect to T, measmd in situ, in a slurry. transmittance at 3394 cm'1 TR = . -1 transmittance at 1172 cm 41 Concentration Temperature Undersaturation V Temperature (a) Figure 3.4 Simulated desupersaturation profiles, (a), of cooling batch crystallization. The desupersaturation profiles 1, 2 and 3 correspond to parabolic, linear and natural cooling profiles, (b), respectively. 42 number of nuclei. Consequently, the average sizes of CSDs resulting from desupersaturation profiles that approach the metastable limit are very small. Such distributions can be overly taxing 0n filtration and drying units, driving up operation costs. Away fiom the metastable limit and close to the solubility line, secondary nucleation (nuclei formed fiom collisions and shear) and growth events are prominent. In this circumstance, due to the lack of driving force, the limited number of nuclei grow slowly resulting in coarser CSDs. Therefore, to meet CSD specifications, the level and the rate of supersaturation should be maintained appropriately. The level of supersaturation corresponding to target product specifications depend on the width of the operating zone defined by the metastable limit and solubility. Therefore, the knowledge of the location of the metastable limit will give engineers a better grasp on system behavior, allowing more effective operations. The metastable limit of a particular system depends on the supersaturation rate and the hydrodynamic conditions of the crystallizer [9] . Since the metastable limit depends on the hydrodynamic conditions of the crystallizer, it is imperative that it be measured under the same conditions that the material will be crystallized. ATR FTIR spectroscopy is a particularly convenient technique that provides the opportunity to measure the metastable limit in situ. The metastable limit can be measm'ed easily as illustrated by experimental data in Figure 3.5. A homogeneous solution at point X can be cooled while accumulating ATR FTIR spectra (T, accumulated every 20 seconds) of the supersaturation profile. At point Y,, the system nucleates spontaneously and rapidly desupersaturates due to 43 8 i i ifi - i C 7.5 "': \ :1- 7 .1 X ,‘. 6.5 .E. En: i i 6 .1 -' 1 . 5,5 .1:- 1 ° Data 3 5 .L Solubility J. C 5 + Metastable limit I 4.5’-~-:--~+~--:--~ 50 55 60 65 70 Temperature (°C) Figure 3.5 The metastable limit of maleic acid, (Y,), measured at a cooling rate of 1.3 °C/min. and a stirring rate of 420 rpm. The result of one experiment is given for illustration purposes. For the given experiment, point X denotes the initial conditions and point Y, denotes spontaneous nucleation. All points, Y,, were extracted from similar experiments. __ transmittance at 3394 cm'1 transmittance at 1172 cm'1 TR 44 excessive nucleation and growth. Thus, point Y, is one point of the periphery of the operating zone for the given system under the implemented cooling and stirring conditions. In this case, the cooling rate was 1.3 °C/min. and the stirring rate was 420 rpm. The cooling rate implemented was the maximum allowed by the capacity of the chiller. A series of experimental points, Y,, constructs the metastable limit for the system under the given operating condition. 3.3.3 Measurement of Supersaturation In Situ The primary impetus for measurement of supersaturation, in situ, is that it allows the estimation of kinetic parameters under hydrodynamic conditions governed by viscous properties of the system, heat and mass transfer properties and mixing patterns. The system kinetics in ttn'n determine the process conditions. Crystallization process models indicate that, to maintain a desupersaturation profile as the one represented by profile number 1 of Figure 3.4, the temperature profile implemented should be parabolic in time [10, 11]. The initial rate of cooling corresponding to a parabolic cooling profile is very low and it increases progressively over the course of the batch time. On the other extreme, a natural cooling profile (uncontrolled cooling) leads to a supersaturation profile (profile number 3) that approaches the metastable limit resulting in extremely fine CSDs. The desupersaturation profile corresponding to the linear cooling profile may follow a course as the one depicted by profile number 2. The challenge in process development of batch crystallization is to ascertain the appropriate controlled cooling profile. The profile cannot be any parabolic or linear profile because the coemcients of the profile depend significantly on the kinetics of 45 the particular system under operating conditions. Therefore, a consistent set of kinetic parameters that adequately describe process behavior is necessary for process development leading to efficient operation of crystallization processes. Models to extract growth kinetics fi'om batch crystallizers and models to analyze batch crystallizers have been developed [12]. However, the use of kinetic models or the application of behavior models for industrial batch crystallizers has been stymied by the unavailability of an in situ measuring device for supersaturation which is the driving force for all crystallization phenomena. In essence, the prerequisite for the determination of coeficients of controlled cooling profiles satisfying product specifications based on system kinetics is the ability to measure supersaturation in situ. In order to demonstrate the importance of in situ measurement of supersaturation for process development and control purposes and the effect of small changes in supersaturation on the product CSDs, the results of three experiments are presented. Results of each of the three experiments consist of a cooling profile, the corresponding desupersaturation profile and the resulting product CSD. The three cooling profiles are given in Figure 3.6. They are identified as parabolic, intermediate parabolic and linear cooling profiles. The corresponding desupersaturation profiles are given in Figures 3.7 (parabolic), 3.8 (intermediate parabolic) and 3.9 (linear). The differences in the course of desupersaturation corresponding to the cooling profiles are clearly evident. The generation of a supersaturation peak as in the case of the desupersaturation profiles corresponding to the intermediate parabolic and 46 64 .1 . . : . . r +Parabolic -5- Intermediate parabolic +Linear Temperature (°C) O 360 720 1080 1440 1800 Time (sec.) Figure 3.6 Experimental cooling profiles. They are identified as parabolic, intermediate parabolic and linear cooling profiles. 47 7.5-; ,-; -:e'- + I, 7 "’ .1 )- Elm 6.5 "' I,” .- , I," 0 Experimental Data 6 -~,I’ — Smoothed Data f , —Solubi].ity ’ """ Metastable Limit 5.5 I If . - - 1' - 4 - 1 56 58 60 62 64 Temperature (°C) Me 3.7 The desupersaturation profile corresponding to the parabolic cooling profile given in figure 3.6. _ transmittance at 3394 cm’1 TR - . _1 transmittance at 1172 cm 48 7.5 . i . . . i . . . : , i, , g . . ' . b ’I D i” & t I," g 36.3"... C}? j 7 .. .,' . 0 4P 1 0 o . a; . ' i a: ’ ’1’ .." t E 6'5 -- 'I’ ”’10 'l' P 1’ 00.1 1 . ’1” o ..,' 0 Experimental Data 6 '2" . 1’ —Smoothed Data ' _./ —Solubil.ity , 7"" """ Metastable L1m' it 56 58 60 62 64 Temperature (°C) Figure 3.8 The desupersatmation profile corresponding to the intermediate cooling profile given in Figure 3.6. _ transmittance at 3394 cm'1 transmittance at 117 2 cm'1 TR 49 7.5 i t : r; . : . - - L x L "I . o 0 ° I, O 7 . x. . Em 6'5 ‘ ’X o o ‘ b I 1 ; I," , - 0 Experimental Data 6 u," . — Smoothed Data f —Solubi1ity /‘ ----- Metastable Limit 55- “‘:r“:‘*‘:“‘1. 56 58 6O 62 64 Temperature (°C) Figure 3.9 The desupersaturation profile corresponding to the linear cooling profile giveninFigin'e3.6. _ transmittance at 3394 cm'1 TR - . -1 transmittance at 1172 cm 50 the linear cooling profiles or the lack thereof as in the case of the desupersaturation profile corresponding to the parabolic cooling profile, essentially determines the characteristics of the final product CSD [1 1]. The magnitude of the supersaturation peak depends on the implemented cooling profile. Thus, we emphasize the importance of extracting kinetic information from in situ measurements in order to determine the appropriate cooling profile. Consequences of implementing the wrong cooling profile can be detrimental. A faster rate of cooling will drive the system closer to the metastable limit resulting in a finer CSD than that specified causing operation problems at filtration and drying steps. On the other hand, unnecessarily slow cooling rates can stagnate the operation due to inadequate driving force resulting in unreasonably long batch times. The experimental cooling profiles in Figure 3.6 were not based on system kinetics, rather they were chosen and implemented based on prior experience with the system to demonstrate the efi‘ect of difl'erences in supersaturation on product CSDs. However, as part of this ongoing efi'ort, the corresponding desupersaturation profiles will be used to extract kinetic parameters that will enable the determination of cooling profiles appropriate for specified CSDs. The subject is taken up in Chapter 5. The mass based product CSDs resulting from desupersaturation profiles corresponding to the three cooling policies are given in Figm'e 3.10. The average size of the CSD resulting from the parabolic cooling profile is about 760 microns. The average size of the CSD resulting fiom the linear cooling profile is about 440 microns. The average size of the product resulting from the parabolic profile is nearly double in comparison to that CSD (% weight) Figure 3.10 51 l I 50 i 4 5 i I . a +Parabolic . f E. "I- Intermediate paraboli 4o -_- 5 ]440 --0-Linear L i' l i . g i 760 . i 1 3o .. f i * i i L g g ’ 20 .. :1 . .5 a,” i ,l.‘ 10 q- '0' ‘ , 3' f If 0' to 0 ' t 200 400 600 800 1000 Size (microns) Mass based product crystal size distributions (CSD). The CSDs 1, 2 and 3 correspond to desupersaturation profiles 1, 2 and 3, respectively. The arrows indicate the mass based average sizes of each size distribution. 52 resulting fiom the linear profile. The changes in CSD characteristics are significant. This evidence underscores the importance of implementing the cooling profile befitting the specified CSD. Thus, it is anticipated that the ability to measure crystallization parameters, especially supersaturation, in situ, will have a considerable impact on the efficiency of process development and control of batch crystallization processes. 3.3.4 Calibrations Calibrations can serve two purposes. First, they demonstrate that ATR FTIR spectroscopy, TR, provides a true reflection of system variables. More importantly, calibrations provide the basis to extract the transient crystal slurry density. The slim-y density is an explicit parameter in most batch crystallization process models [12, 13]. The calibrations of TR for aqueous maleic acid with respect to temperature and concentration (% w/w maleic acid in water) are given in Figure 3.11. The plot consists of a three parameter grid comprising TR, temperatme and concentration on which the experimental solubility has been superimposed. Each point where the solubility line crosses the concentration lines indicates the solubility of maleic acid in water with respect to the solution concentration which is the more conventional way of representing solubility. This data is within 3% of the solubility data given in [9] . The accuracy of the current experimental data confirms that ATR FTIR spectroscopy is well suited to measure crystallization parameters. Secondary nucleation depends on the slurry density. Therefore, slurry density is an explicit parameter in most batch crystallization process models. It is important to be able to measure the transient slurry density to 53 Ill- - ) 8--r--g----;-,-- "T‘"'.%(w/w) .. . . Solubihty v 165% i 64% 63% j- 62% ‘. 61% j_ 60% 70 Temperature (°C) Figure 3.11 Calibrations of the transmittance ratio, T3, with respect to temperatm'e and maleic acid concentration (% w/w maleic acid in water). The constant concentration lines were extracted from the given experimental data. _ transmittance at 3394 cm’1 transmittance at 117 2 cm’1 TR 54 extract kinetic parameters and to accurately emulate system behavior. The equation; C=(a1+a2xT+a3xT2)(bz+b2xTR+b3xT3) (3.2) where T is the temperature, 01=l.079, a2=-1.6989x10“3, a3=5.7812x10“, b1=40.39, b2=5.4039 and b3=-2.77O4x10'°1, gives the solution concentration of maleic acid in water at any point in the three parameter grid of Figure 3.11. Using this equation, the mass balance, the initial conditions and the batch volume, the transient slurry density can be easily calculated. There is a noticeable increase in the noise level of the calibration data compared to the solubility data. This is due to the accumulation of air bubbles on the sensing head of the internal reflection element of the immersion probe. Air bubbles were a problem in the 2L crystallizer used for these experiments since its built-in baflle design is not adequate to completely avoid vortex formation. In the presence of a slurry, however, the crystals of the slurry inhibit the accumulation of air bubbles on the sensing head. Consequently, the noise level is dramatically reduced. This is clearly demonstrated by the low noise level in the solubility data that were measured in slurries. It is anticipated that air bubbles will not cause significant problems since the true measuring environment for crystallization parameters is a slurry. 3.4 Conclusions The primary objective of this effort was to investigate the feasibility of ATR FTIR spectroscopy for in situ measurement of crystallization 55 parameters, especially supersaturation. To this end, the data have not only demonstrated that in situ ATR FTIR spectroscopy is well suited to measure supersaturation, the data have also demonstrated that ATR FTIR spectroscopy is sensitive enough to measure small changes in supersaturation that lead to significant changes in product CSDs. It is anticipated that the technique will be useful in process development and control of batch crystallization, providing for improved product quality/reproducibility and reduced energy utilization and cost of operation. 3.5 References 1. Zimmerman, H. K., Chem. Rev., 51, (1952), 25. 2. Liszi, I., Halasz, S., Bodor, B., in: “Industrial Crystallization 87: Process Technology Proceedings, 6”, Ed. Nyvlt, J. ; Zacek, S., Elsevier, New York, USA., (1989), 399. 3. Ymamoto, H., Sudo, S., Yano, M., Harano, Y., in: “Industrial Crystallization 87 : Process Technology Proceedings, 6”, Ed. Nyvlt, J .; Zacek, S., Elsevier, New York, USA, (1989), 403. 4. Myerson, A S., Rush, S., Schork, F. J ., Johnson, J. L., in: “Industrial Crystallization 87: Process Technology Proceedings, 6”, Ed. Nyvlt, J .; Zacek, S., Elsevier, New York, USA, (1989), 407. 5. Miller, S., Rawlings, J. B., AIChE J, 40:8, (1994), 1312. 6. Nyvlt, J ., Karel, M.; Pisarik, S., Cryst. Res. Technol., 29, (1994), 409. 7. Mirabella, F. M., Ed., “Internal Reflection Spectroscopy”, Dekker, New York, USA., (1993). 8. Dunuwila, D. D., Caroll 11, L. B.; Berglund, K. A., J. Cryst. Growth, 137, (1994), 561. 9. Mullin, J. W., “Crystallization”, Butterworth-Heinemann, Oxford, England, (1993). 10. Mullin, J. W., Nyvlt, J ., Chem. Eng. Sci., 26, (1971), 369. 11. Jones, A. G., Mullin, J. W., Chem. Eng. Sci., 29, (1974), 105. 12. Tavare, N. S., Separation and Purification Methods, 22:2, (1993), 93. 13. Randolph, A. D., Larson, M. A, “Theory of Particulate Processes”, Academic Press, New York, USA., (1988). * CHAPTER4 IDENTIFICATION OF FTIR SPECTRAL FEATURES RELATED TO SOLUTION STRUCTURE FOR UTILIZATION IN NIEASUREMENT OF CRYSTALLIZATION PHENONIENA 4.1 Background Carboxylic acids are generally characterized by vibrational modes of the carboxyl group [1]. Of the six vibrational modes of the carboxyl group, the v(C=C), v(C-O) and v(O-H) stretching vibrations and the 5(O-H) in-plane deformation vibration produce prominent fundamental IR absorbtion bands in the 1100-3000 cm" region of the mid-IR spectrum. The nomenclature is discussed in [2]. The distinctive structure of the carboxyl group facilitates association of carboxylic acids through hydrogen bonding [1]. Association of carboxylic acids is illustrated in Figure 4.1. Formation of these dimers lowers the force constant of the v(C=C) and the v(O-H) vibrations and the frequency of the vibrations [3, 4]. The large decrease in fiequency as a result of the transition fiom the monomeric to the associated form is evidence for the exceptional strength of hydrogen bonds. Hydrogen bonds, due to their strength, are largely responsible for the solution and crystal ' Dunuwila, D. D., Berglund, K. A., Submitted to Trans. I. ChemE., (1995). 56 Figure 4.1 57 /O----H—O\ R—C /C —R \O—H-n-O/ Self association of carboxylic acids via H-bonding. 58 structure of carboxylic acids. Consequently, any change in solution structure due to reactions, changes in composition, introduction of new species and temperature is reflected by spectral features produced by molecular vibrations affected by hydrogen bonding. Carboxylic acids can form complexes with polar solvents via hydrogen bonds [1, 5, 6, 7]. Depending on the proton donor/acceptor strengths of polar solvents, carboxylic acids can form hydrogen bonds either by the participation of the carbonyl oxygen or the hydroxy group. However, intermolecular hydrogen bonding of dissimilar species is significantly weak compared to that of the same species. 4.1.1 Identification of the Vibrations of the Carboxyl Group The v(O-H) Stretching Vibration: The v(O-H) vibration for monomeric carboxylic acids is observed around 3500-3570 cm'1 [1, 8, 9]. The monomeric form occurs only in the vapor phase and in dilute solutions of nonpolar solvents. The broadened v(O-H) band of the associated form is observed over the range 2700-3100 cm‘1 [1, 3, 4, 10, 11]. The associated form, either dimeric or polymeric [9, 12] , occurs almost exclusively in solid state, liquid state and in concentrated solutions. The shift of the v(O-H) band to lower fi'equencies upon association is characteristic of carboxylic acids. In addition, the broad band generally appears as a convoluted band made up of several bands. A simple interpretation is that the overlapping bands are due to a number of hydrogen bonds of varying lengths resulting fiom the complexity of associations [1]. An alternate interpretation is that this is due to coupling of the v(O-H) fundamental with several other low 59 frequency vibrational modes of the associated form [1, 13]. Other explanations are available [10]. The v(C=O) Stretching Vibration: The v(C=O) vibration for monomeric saturated aliphatic carboxylic acids is observed around 17 80:15 cm‘1 [1, 3, 9]. The v(C=O) vibration for the associated form of saturated monocarboxylic acids (liquid and solid state) is found around 1710i10 cm'1 [1, 3, 9, 10, 11] demonstrating the shift to lower frequencies upon association. Both the monomer (around 1760 cm") and the dimer (around 1710 cm“) are observed in the case of many carboxylic acids in nonpolar solvents [1, 14]. The v(C-O) Stretching and the 5(O-H) In-Plane Deformation Vibrations: Coupling of the v(C-O) stretching and the 5(O-H) in-plane deformation vibrations occurring in the plane of the ring formed by the association of carboxylic acids result in two bands around 1420i20 cm'1 and 1300:]:15 cm'1 [1, 10, 11, 15]. The band at 1420 is assigned to v(C-O) and the band at 1300 is assigned to 5(O-H) [3]. These bands are reported to appear only under conditions that promote association via hydrogen bonding [1, 15]. These two bands are reported to shift in the opposite direction to that of v(C=O) and v(O-H) upon association [3, 9]. 4.2 Materials and Methods All details regarding the experimental setup and instrumentation are given in Chapter 3 and [16]. The protocols for the accumulation of solubility 60 data and data corresponding to the constant concentration lines are outlined below. Further details are given in Chapter 3 and [16]. 4.2.1 Measurement of Solubility in Slurries Solubility was measured in a slurry, in situ, at temperatures ranging from 30°C to 70°C. At each temperature, ten spectra were scanned. Therefore, the plotted data represent the average of ten measurements. An equilibration time of about one hour was allowed for the slurry at each temperature setting. A slurry equilibration time of one hour was determined to be sufficient by monitoring the system approaching equilibrium both fi‘om supersaturation and undersaturation. 4.2.2 Construction of the Calibration Curves Appropriate amounts (weight) of maleic acid and distilled water were placed in a 2 L crystallizer and heated to a homogeneous solution. The solution was cooled while accumulating solution ATR FTIR spectra, in situ, every 2 minutes until it spontaneously nucleated. A series of such experiments ranging fi'om 50 to 65 percent maleic acid by weight was conducted. 4.2.3 Data Processing Deconvolution: The spectral region between 2750-3700 cm“1 was decon- voluted as follows. The given region was isolated fi‘om rest of the spectrum. The baseline of the absorbtion spectra was fixed at 37 00 cm'1 (that is, absorbance at 3700 cm'1 was set to zero). Spectra were normalized at 3490 cm". All spectral profiles were fitted to a function consisting of six Gaussian components using the “leastsq” least squares minimization function available in the professional version of the MATLAB optimization toolbox. 61 The six-component Gaussian function provided the best fit for the spectral profile. The contribution of each component to the spectral profile was generated using the adjustable parameters (peak position and standard deviation (peak width at half maximum» of the corresponding Gaussian distribution. Derivative Spectra: The derivative spectra of ATR FTIR absorbtion spectra were obtained. The peak positions of all the identified maleic acid vibrational modes and the useful peak intensities (intensities used to compute PIR1 and PIRZ) were extracted from the derivative spectra and tabulated. This otherwise tedious task was simplified by using the user programmable OBEY programming utility available in the Perkin-Elmer IRDM software. 4.3 Results and Discussion As discussed earlier, the primary focus of this paper was to investigate the methods available to identify and isolate the subtle features of IR spectra, that are susceptible to changes in solution structure, for utilization in measurement of crystallization parameters. Changes in molecular vibrational structure are manifested in spectral features such as shifts in peak positions and peak widths that may not be readily apparent [1] . Both derivative spectroscopy [17] and deconvolution [17] of overlapping bands can be used to isolate peak position and peak width shifts of IR absorbtion bands arising from reorganizations in solution structure. Derivative spectroscopy, which is a standard menu option in FTIR software, is the more straightforward method of the two and thus will be discussed 62 first. Although, some FTIR instrument manufacturers include a deconvolution algorithm as a standard menu option, these algorithms are designed only for peak enhancement and thus cannot resolve highly convoluted broad bands. Spectral data can be cluttered by interferences and artifacts that appear in the form of polynomial shapes such as sloping lines or quadratic curves [17]. The derivative of the spectrum, which is one order higher than the polynomial shape, resolves some of the finer details concealed by the polynomial shapes. For example, the first derivative removes sloping lines and the second derivative removes quadratic curves and adds clarity to the spectrum. As such, derivative spectra can reveal a host of information otherwise hidden or not readily apparent in the original spectra and can be used to extract both qualitative and quantitative information. Figure 4.2 illustrates the derivative curves of a Gaussian band. The first and the third derivative curves give the precise peak position while the second and the forth derivative curves generate the inverted and the original curves, respectively. The illustrations are not intended to demonstrate the particular advantages of derivative spectroscopy, rather they demonstrate the results of taking the derivative of an ideal hand up to the fourth derivative. Their utility in analyzing real spectra will be apparent shortly. Figure 4.3 gives the ATR FTIR spectra (lower frequency window) of saturated aqueous maleic acid at various temperatm‘es. Band assignments are indicated by the arrows [1, 3, 9, 10, 11, 14, 15]. These spectra are used to illustrate the advantages of derivative spectroscopy. 63 ---- Gaussian ---- Gaussian — Second derivative — First derivative A "fVWVfijw- A A Vii v v v va v “"‘l‘l‘“"“““ ---- Gaussian ---- Gaussian — Fourth derivative [‘1‘ ‘ ‘ ‘ F‘ i ‘ ‘ ‘ 1 d ‘ ‘ I ‘ ‘ ‘ I A A A A A A A A A A A A A A — Third derivative 1650 1600 1650 1550 1600 1550 Wavenumber (ch) Figure 4.2 First, second, third and fourth derivatives of a Gaussian peak. (172 - ‘ 4. 4+4 . .. i. .. . E. .1». 5. -. . t. .. +3075 °C (164‘# +50.25 °C D + O 0.56 . 65.00 C a) o . g 0.48 - 4: . H 8 04 ' .Q ' ' <2 (132 i / (L24 0.16’ll A: ‘ ‘ 1:1411: L 1700 1600 1500 1400 1300 1200 Wavenumber (cm'1 ) Figm'e 4.3 In situ, slurry ATR FTIR spectra of aqueous maleic acid at sattn'ation in the low frequency region. 65 Consider the v(C=O) peak that occurs in the vicinity of 1700 cm". It appears to be a single peak devoid of any other feature such as shoulders. However, the first derivative profile given in Figure 4.4 reveals information plausibly hidden by an artifact in the form of a polynomial shape. The most striking feature is the enhancement of a shoulder on the high frequency side of the v(C=O) peak that is manifested in the form of two peaks in the positive component of the derivative profile. A parameter can be defined based on a ratio (PIR1) between the derivative absorbance peak intensity at 1740 cm'1 to that at 1720 cm’1 for the measurement of crystallization phenomena. A higher gain in PIR (peak intensity ratio) can be achieved by using an appropriate baseline as the one denoted by BLl at 0.005. The solubility and the concentration lines of maleic acid with respect to PIR1 measured in situ are given in Figure 4.5. This parameter is clearly a function of temperature and independent of concentration. Therefore, the concentration lines coincide with the solubility line. Consequently, this parameter is not suitable for the measurement of supersaturation of aqueous maleic acid. However, the results are presented to demonstrate the power of derivative spectroscopy in bringing out spectral features that may be used to measure crystallization parameters of many other systems. Alternatively, a second parameter can be defined using the features of the first derivative of the v(C=O) mode. The defined parameter is the ratio (PIR2) between the derivative absorbance peak intensity at 1740 cm‘1 to that at 1692 cm'1 with the baseline adjusted to -0.012 (BL2) as indicated in Figm'e 4.4. Essentially, this parameter reflects the relative changes 66 0.015P-v-ivv-l ............. l t A +3075 00 001 . " -B-50.25°C. Q) ' . ‘2 8 ’ " +65.00°C é EL 1 T A\’ _ o 0.005 ’ I \\ m ’ \ e + _ 1 e 0.. 1740 cm'1 \‘ 692 cm' . .2 : \\ . 45 » \\ , .> ' _1\“ / 5'0-005'1 1720 cm .. Z/ . r: -0.01 13L 2 . 17 80 1760 1740 17 20 1700 1680 Wavenumber (cm'l) Figure 4.4 First derivative profiles of the v(C=O) vibrational mode fiom Figtn'e 4.3. BL1=0.005 and BL2=-0.012 were the baselines used for P112, and Pm.” respectively. PIR,=(DA at 1740 cm‘1 - a)/(DA at 1720 cm’1 - a) Pm2=(DA at 1740 cm'1 + b)/(DA at 1692 cm’1 + b) a=0.005 and b=0.012 are the baseline adjustments DA=derivative absorbance 4.5 PIRl 00 2.5 ‘f l 3 .5 ’ -O- Solubilityl 50% 53% 56% 59% 62% 65% ZBDODO 2t 1.5 30 35 40 45 50 55 60 65 70 Temperature (°C) Figtn'e 4.5 Solubility and concentration lines of maleic acid with respect to PIR1 measured in situ. In addition, solubility was measured in a slurry. PIR,=(DA at 1740 cm‘1 - a)/(DA at 1720 cm'1 - a) DA=derivative absorbance a=0.005 is the baseline adjustment 68 between the high fi'equency side of the v(C=O) peak and that of the low frequency side. The solubility and the concentration lines of maleic acid with respect to PIR2 measured in situ are given in Figure 4.6. PIRg, unlike PD?” is a function of concentration and only slightly dependent on temperature as demonstrated by slightly sloping concentration lines. The fact that P11?2 is a strong ftmction of concentration renders it a highly suitable candidate for measurement of supersaturation as demonstrated in Figure 4.6. Previously [16, 18] , it has been established that band ratioing techniques such as PIRs or area intensity ratios are the most appropriate for quantitative measurements. This is because ratioing gives rise to an internal standard that efi'ectively eliminates errors due to instrumental drifts. Instrumental drifts usually occur due to energy fluctuations that arise as a result of radiation source instabilities and alignment problems of optical coupling attachments. In many systems, it may be diflicult to identify features that can be used for ratioing in the original spectra. Features of infrared spectra of crystallization systems can be masked by the presence of various species such as impurities, byproducts and additives that have strongly absorbing vibrational modes making the identification of features suitable for ratioing dificult. In addition, most solvents have characteristic IR spectra. As demonstrated above, derivative spectroscopy can be successfully used to bring out features that are well suited for ratioing. In this case, the v(C=O) peak used for demonstration is devoid of any identifiable feature in the original spectra. Nevertheless, the features that the first derivative profile reveals are quite remarkable. This is ample 69 12 "Hi --1‘....: i. -+,, :'j1: r ’ o 50% . D 53% ‘~‘ .5523: ’ o 56% - a 10 4. b 59% " t. ’ a 62% . . a 65% .,::..:_-::_.B.31': :: 8 ._ +Solubility| " - -- " _ on" , ‘ H m _ A 6 .. 1,. ’ 4 4 . . 'l.‘ ‘ ‘- 30 35 40 45 50 55 60 65 70 Temperature (°C) Figure 4.6 Solubility and concentration lines of maleic acid with respect to PH?2 measured in situ. In addition, solubility was measured in a slurry. Pm2=(DA at 1740 cm" + b)/(DA at 1692 cm‘1 + b) DA=derivative absorbance b=0.012 is the baseline adjustment 70 evidence for the power of derivative spectroscopy and in general for the far reaching utility of ATR FTIR spectroscopy for use in measuring and understanding crystallization phenomena. As illustrated in Figm'e 4.4, the second derivative profile of a peak generates the inverted peak. It can be used to isolate the individual components of a convoluted band. Consider the v(C-O) peak at 1430 cm‘1 that is accompanied by three other shoulder peaks that may be due to vibrational mode coupling or overtones. The second derivative can be used to isolate the v(C-O) peak from the shoulder peaks. The inverted second derivative of the v(C-O) stretch at saturation is given in Figme 4.7 . The relative changes of this peak in the original spectrum are identifiable. However, the same changes are more recognizable and convincing in the second derivative profile which is the isolated peak. The changes can readily be used for measurement of crystallization parameters. For example, a PIR between peak intensities at 1430 cm'1 and 1442 cm'1 is suitable. Parameters discussed thus far for quantitative measurements were based on hand ratioing techniques. Although they are highly suitable for quantitative measurements their relationship to solution structure may be ambiguous. A parameter that is suitable for quantitative measurements and capable of providing some insight to the solution structure of crystallizing systems is the peak shift of numerous vibrational modes. Just as H-bonding is responsible for causing the transition of monomeric carboxylic molecules in dilute solutions to associated forms in more concentrated solutions [1, 19] , they are responsible for the crystal structure 71 (105 - -- -5 -- -- 5- -, -5 fjgv- e- -- -1 -- { +30% °C j _ —B—40.50 °C , +5050 °C A t 0.04 .. +6000 °C . .. r +6500 °C . r . g 0.03 -. 0. “Q I 1 H 1 O r 3 002 -’ / J <1 ' , /,~ 0-01 ~ 1430cm'1 ‘- .. 1442 cm'1 1 » y’)- 0 . :2} AAAAAAAAAAAAAAAAAAAAAA ‘5 1460 1450 1440 1430 1420 1410 Wavenumber (cm'l) Figure 4.7 Isolated second derivative profiles of the v(C=O) vibrational mode of aqueous maleic acid at saturation. 72 and the preceding solution structure in supersaturated solutions. As the forces such as H-bonds that are responsible for the organization of molecules in supersaturated solutions gain in strength, the force constants of the vibrational modes adjacent to the site affected by H-bonding are reduced. The consequent reduction in vibrational frequency is reflected in the IR spectrum [3, 4]. Due to the presence of electronic phenomena such as inductive effects and conjugation [1] , peripheral vibrational modes may be affected and the resulting structure is reflected in the IR spectrum. Derivative spectroscopy is a convenient method for extracting peak shift information. The same information can be obtained from the original spectrum. However, the shifts are more discernible in derivative spectra (lst and 3rd derivative). Consider the 8(O-H) in-plane deformation vibration of maleic acid given in Figure 4.3. The red shift of the peak upon increasing saturation temperature conditions is evident. The same shift is more convincingly demonstrated by the derivative profiles of the peaks in Figure 4.8. The parameter Av denotes the total shift of the peak in going from the solubility condition at 30.75 °C to that at 65 °C. The peak shift information was extracted using the user programmable OBEY programming utility available in the Perkin-Elmer IRDM software. A discussion of the use of maleic acid vibrational modes for the measurement of crystallization parameters of aqueous maleic acid and some insight to the organization of maleic acid in supersaturated solutions follows. The covalent bond lengths and the interbond angles for maleic acid derived fiom the crystal structure are shown in Figure 4.9 [20]. The IR peak 73 0.0053 " 0.0027 0 -0.0027 Derivative Absorbance O I +3075 °C '0'0053 , —a—50.25 °C *‘ 74 I +6500 °C ‘ / -0.008" 1260 1248 1237 1225 1213 1202 Wavenumber (cm'l) Figme 4.8 First derivative profiles of the 5(0-H) in-plane deformation vibrational mode from Figure 4.3. 1.218 \~ 118.9" 1.04 1.337 113° — - 7 131.6° 1.488 121.4° .304 110° 171° 0.91 74 1.0 , ® 116° 128.2° 1.475 1 3 108° 0.99 125.1 ° 6 ' 122.6° 1.222 ® 110° un— g, 1.595 figme 4.9 Bond distances and angles for the maleic acid molecule. Bond (fistancesrnnagfivenuhillaundtineauufiku;areitahcinadfbrcflarfimn 75 positions of the numerous vibrational modes of crystalline maleic acid are as follows [1, 9, 21]. The v(C=O) stretching mode appears at 1705 cm". The v(C—O) stretching mode appears at 1437 cm“. The v(C=C) stretching mode appears at 1635 cm". The v(O-H) stretching vibration appears over the range 2500 - 3000 cm". This band is modified by the v(C-H) stretching hands that occur over the 2800 - 3100 cm‘1 range. The 6(O-H) in-plane deformation appears at 1272 cm". Evidence suggests that the folded structure of the maleic acid molecule in the crystalline state - due to intramolecular H-bonding - stays intact upon dissolution in concentrated solutions (0.3M) [5]. Therefore, it is plausible to expect a high level of similarity between the crystalline state IR spectra and the solution state IR spectra of maleic acid. A close comparison of the two confirms this rationalization. Upon supersaturation of aqueous maleic acid - that is when a solution is cooled and as a consequence it approaches the onset of nucleation - it is also expected that the solution state peaks will approach the solid state peak positions, given the similarity between the solid and solution state molecular structure. An absolute convergence is unlikely due to the higher level of organization in the crystal lattice. The following paragraphs examine the events that are expected to accompany the convergence of solid and solution state structures in terms of peak shifts and the utility of these shifts toward measurement of crystallization parameters. 76 Solubility and concentration lines of aqueous maleic acid with respect to the peak shift of the v(C=O) vibration measured in situ are shown in Figure 4.10. Solubility, in addition to being measured in situ, was measured in a slurry. When solutions at all concentrations are cooled, the v(C=O) stretching vibration red shifts to lower frequencies. This is clearly indicative of the strengthening H-bond. Upon cooling, H—bonds gains in strength. Consequently, the strength of the adi acent C=O bond or the force constant of the v(C=O) stretching vibration is reduced. This phenomenon is manifestedinthe IRspectrumas ared shiftinthev(C=O)peak as shown in Figure 4.10. Solubility and concentration lines of aqueous maleic acid with respect to the peak shiit of the v(C-O) vibration measured in situ are shown in Figure 4.11. Upon cooling, \mlike the v(C=O) peak, the v(C-O) peak demonstrates a blue shift. This is consistent with previous studies on association of carboxylic acids [3, 9]. The complexity of electronic effects such as inductive forces and conjugation that are affected by H—bonding may be responsible for the blue shift of the v(C-O) peak. Solubility and concentration lines of aqueous maleic acid with respect to the peak shift of the v(C=C) vibration measured in situ are shown in Figure 4.12. When solutions at all concentrations are cooled, the v(C=C) stretching vibration red shifts to lower frequencies. Since most studies on association have been done using saturated carboxylic acids, peak shifts of 77 0 -__-,----;,---;----;,,,-;----;,--T+--,- * 0 50% ‘ r U 54% 1 o 58% -2 .. 0.“, ‘. A 62% .. El . B 65% J a " ;_. - + Solubility , 5 .ui" o 5. '4 ' l fl ' 1706 cm'l ‘ 'M I- o -1 1 8 .~ 1708 cm . O-t -6 .b A ‘h ‘ -8 a. “nil" un- “flfi 30 35 40 45 50 55 60 65 70 Temperature (°C) Figure 4. 10 Solubility and concentration lines of maleic acid measured with respect to the peak shift of the v(C=O) vibrational mode, in situ. In addition, solubility was measured in a 31m. Absolute peak positions at two conditions are shown for reference purposes. 78 14 :r -; .4, - : :vm-n—r t t t o 50% fi . 12 .. n 54% B -==.: .. : o 58% ,3 . ' A 62% ‘ “. 1° '3 a 65% n ‘ Ba 1‘ A Z+Solubility o ' 1 8 3.. ° .. t: A 4 E ED ,0. 3 LE 6 L o n ° ° 1427 cm'1j_ s. m t D" Q o ‘ l: '33 ’ . ° , q) 4.. 1. DH r O ‘ ‘ ; o '. n n ; 2.. o .. .. . '3 1429cm'l : 0 1" El ° 1 -2 111. 30 35 40 45 50 55 60 65 70 Temperature (°C) Me 4.11 Solubility and concentration lines of maleic acid measured with respect to the peak shift of the v(C-O) vibrational mode, in situ. In addition, solubility was measm'ed in a slurry. Absolute peak positions at two conditions are shown for reference purposes. 79 14 4 C 9 12. . .. 103 I. ’5‘ o -°°° 1 v 8 -- 1627 cm’1 4 :1 - - 9 CD 6 “ l3"’ 1:5: 4 1 9 . m . o 50% I 4- n 54% v °" ° 58% . 2 d; A 62% :_ -1 B 65% . <———1630.5 cm _ Solubility~ 0 - 3O 35 4O 45 5O 55 60 65 70 Temperature (°C) Figure 4.12 Solubility and concentration lines of maleic acid measm'ed with respect to the peak shift of the v(C=C) vibrational mode, in situ. In addition, solubility was measured in a slrn'ry. Absolute peak positions at two conditions are shown for reference purposes. 80 the v(C=C) vibration are not documented. Nevertheless, this data suggest that the strength of the C=C bond or the force constant of the v(C=C) stretching vibration is reduced upon approaching the onset of nucleation. Solubility and concentration lines of aqueous maleic acid with respect to the peak shift of the 5(O-H) in-plane deformation vibration measured in situ are shown in Figrn-e 4.13. The behavior of the 5(O-H) peak shift upon approaching conditions that favor a higher degree of association, in this case lower temperatures, is similar to that of the v(C-O) peak as has been documented [3, 9]. The v(C-O) and the 5(O-H) modes are mutually coupled and blue shifts upon association via H-bonding [3, 9]. The peak shifts of v(C=O), v(C-O), v(C=C) and 5(O-H) vibrations were isolated using derivative spectroscopy. A common feature of these bands is that they have relatively narrow band widths that give rise to informative derivative profiles. However, the derivative profiles of the highly convoluted, broad v(O-H) stretching vibration region are not conducive to meaningful analysis. In the case of aqueous maleic acid, the v(O-H) stretching region constitutes of contributions from both water and acid. In addition, the v(C- H) stretching mode coincide in this region [9]. The relative changes in this region are shown in Figure 4.14. However, subtle features such as peak shifts are not quite detectable. Isolation of the contributing components via deconvolution of the broad, convoluted band can reveal the concealed information. 81 70 7 7 7 r: ......... i ......... i ........ r3 , at“ 13 Z 0 50% :9 f “ " 1 60,; :1 54% , . A/ 1 ; ° 58% . . ‘ : ‘5 62% ' ' a 50.. B 65% .7, -- a I O Solubility 4.’ G .1 . g 40 : . o" 1221 cm'13 6‘5 30.: 3. ":3 i : Q) ., . Dd . ' 'l- J 20 7.: . r 455: 1' 10 .E .‘7 - 1232 cm'1 : 0 1 ' ‘ 30 35 40 45 50 55 60 65 70 Temperature (°C) Figure 4. 13 Solubility and concentration lines of maleic acid measured with respect to the peak shift of the 5(O-H) vibrational mode, in situ. In addition, solubility was measured in a drum. Absolute peak positions at two conditions are shown for reference purposes. 82 0.4 ..... ..... I —e—30.75°C . : ‘ —B—50.25 °C 1 03 .. ‘ +65.00°C ._ 0.2 «i ,7 ' ‘5 .. \ 1 " "x . v ' \V?’ \ ‘ v(O-H)water ' ' ' Absorbance v(C-H) acid ' ' v(O-H) acid 3300 3000 2700 2400 Wavenumber (cm'1 ) Figure 4.14 In situ, slrn'ry ATR FTIR spectra of aqueous maleic acid at saturation in the high frequency region. 83 Deconvolution algorithms are available in some FTIR instrument software packages. However, they are designed to enhance existing peaks. They are not adequate for resolving highly convoluted, broad bands. An alternative is to fit the spectral profile in this region to a multicomponent Gaussian function using a least squares minimization algorithm. A rule of thumb that is generally applicable in utilizing this technique is to use the least number of Gaussian components that provide the best fit for the spectral profile. Consider the spectrum at solubility at 30.7 5 °C over the 2700 - 3700 cm‘1 range (Figure 4.14). The normalized and baseline adjusted spectrum was fit to a function consisting of six Gaussian components. The results are given in Figme 4.15. Due to the closeness of the fit, the spectrum and the six component function used for the fit are not distinctly identifiable in the figm'e. The six isolated components are also shown. The components toward the high fi'equency side of the band can be loosely assigned to the v(O-H) stretching bands of water. Conversely, the components toward the low frequency side of the band can be loosely assigned to the v(O-H) stretching bands of maleic acid. The assignments are not meant to be definitive because the contributions from the v(C-H) stretching vibration of maleic acid to both v(O-H) stretching band intensities cannot be discounted. However, the general trends of the components can be used within the scope of this investigation. For example, it is reasonable to associate most of the peak shifts observed in individual components to the influence of H-bonding on the v(O-H) stretching vibrations. As a result of 84 14 1 1 1 1 1 1 . i I I I I I I U 1 Tfi' V I j I V V Y I’ Y Y‘rj V v(O-H) water 1.2 .L I I .1. 3150 cm" Absorbance 3800 3600 3400 3200 3000 2800 2600 Wavenumber (cm '1 ) Figure 4.15 Gaussian components of the v(O—H) region Due to the closeness of the fit the original spectral profile (normalized and baseline corrected spectrum at saturation at 30.7 5 °C from Figure 4.14) and the six component Gaussian fimction used for the fit are superimposed. 85 the transition from the monomeric to the associated form, a large decrease in frequency of the v(O-H) vibration of carboxylic acids is observed due to the exceptional strength of hydrogen bonds [3, 4]. The same trend is expected to continue in highly concentrated solutions where strengthening H-bonds play a critical role in the organization of carboxylic acid molecules. Consider the component identified at 3150 cm'1 in Figme 4.15. Its peak shift is shown in Figm'e 4.16. As discussed in the previous paragraph, the peak shift was associated with the v(O-H) stretching vibration of maleic acid. The peak shift of the low frequency side of the v(O-H) band is clearly demonstrated by the considered component whereas it is not evident in the original spectra. Solubility and concentration lines of aqueous maleic acid with respect to the peak shift of the v(O-H) vibration measured in situ are shown in Figm‘e 4.17. When solutions at all concentrations are cooled, the v(O-H) stretching vibration is expected to demonstrate a red shift toward lower frequencies under the influence of strengthening H-bonds. The concentration lines appear to have slight negative gradients as expected. However, due to the complexity of this region it is not possible to provide a definitive interpretation. Overall, all the peaks characteristic of maleic acid shift in the direction that suggests a greater degree of association of the molecules (via H-bonding) upon cooling. These trends observed in supersaturated solutions in approaching the onset of nucleation, is consistent with the trends observed in the transformation from the monomeric form to the associated 86 (13 77-9 .7..:. -7 77 --49-- -9 -- -9 1 —9— 30.75 °C 3 0.7 —a— 50.25 °C -j + 60.00 °C 0.6 0.5 0.4 Absorbance 0.3 0.2 0.1 3383 3267 3150 3033 2917 2800 Wavenumber (cm '1 ) Figure 4.16 Peak shift of the component near 3150 crn'l from Figure 4.15. 87 100 7....:....:....:....:.--.:....:....9-. I o 50% v ~. , D 53% ‘ .L ° 56% .. 80 . A 59% B ” -- == , B 62% . a El 65% , 5 60 .. +Solubility ._ " . .. 5:3 . .2 1 U) .14 g 40 '1' o .. 0.. r 20 "' 4p 0.0 0 #1--:----:-1--:---- 3O 35 40 45 50 55 60 65 70 Temperature (°C) Figure 4.17 Solubility and concentration lines of maleic acid measured with respect to the peak shift of the v(O-H) vibrational mode, in situ. Inaddfinursdubflfivaunmmamnednnashury 88 form in more dilute solutions. In addition, all peaks approach the peak positions of the solid state spectrum. While recognizing that this evidence alone is not sufficient to provide conclusive remarks regarding the similarity of the solid state and the solution (supersaturated) state molecular structure, it is not unreasonable to suggest that the molecular structure in the two states are most likely similar considering the convergence of peak positions and the evidence [5] for the folded configuration of maleic acid in solution. A concrete analysis on the nature of the solution state structure of aqueous maleic acid in supersaturated solutions is beyond the scope of this investigation. This analysis is a prelude to a more in-depth study that will involve both Raman and NMR spectroscopy in addition to FTIR spectroscopy. The different techniques can reveal complementary information. Frn'ther, utilization of the deutero analogues of both water and maleic acid in experiments similar to ones discussed above can provide a host of confirmatory information [10, 11, 22] . The deuteration of a particular bond results in a significant red shift in its vibrational frequencies due to the heavy atom effect. For example, the v(O-D) vibration of the deuterated acetic acid is observed at 2299 cm'1 in comparison to the v(O-H) vibration of acetic acid that occurs at 3140 cm‘1 [10]. The 5(O-D) vibration is observed at 1046 cm‘1 compared to the 5(O-H) vibration at 1294 cm“. Consequently, experimentation with selectively deuterated maleic acid and deuterated water in judiciously chosen combinations can help resolve many of the ambiguities of the current analysis. For example, use of maleic acid in 89 heavy water will clear the v(O-H) stretching region of any contribution from the v(O—H) vibration of water thus making it much easier to identify dynamics of the v(O-H) vibration of maleic acid. Thus far, a discussion of the utility of the peak shifts of maleic acid vibrational modes for the measurement of crystallization parameters of aqueous maleic acid, which incidentally is the primary objective of this investigation, has been neglected. The discussion has been centered on providing some insight to the organization of maleic acid in supersaturated solutions. However, maleic acid vibrational modes suitable for measurement of supersaturation and solubility are quite evident. The v(C=C) mode (Figure 4.12), the 5(O-H) mode (Figm'e 4.13) and the v(O-H) mode (Figure 4.17), that demonstrate a strong dependence on concentration, are the ones suitable for this purpose. The other modes that are independent of concentration, although reflective of the solution structure, are clearly not useful for quantitative measurements. Between the band ratioing techniques investigated in Chapters 1 and 2 [16, 18] and discussed herein and the peak shifts of numerous vibrational modes, it is anticipated that it will not be a diflicult task to identify a parameter appropriate for measurement of crystallization phenomena of most inorganic and organic systems. As indicated above, features of infrared spectra otherwise suitable for quantitative measurements of many other crystallization systems can be masked by the presence of various species such as impurities, byproducts, additives and solvents that have, strongly absorbing vibrational modes. Upon encountering such a system, 90 the techniques discussed herein can be used to help enhance features toward identifying a parameter suitable for measurement of solubility and supersaturation. In the case of maleic acid many options that are equally efl'ective are available. The solubility measured with respect to each of the options discussed above is shown in Figure 4.18. The precision of the data is quite reasonable considering the diversity of the parameters used to extract them. In Chapter 1, an inconsistency between the solubility data extracted from two different ratios (RT1 and RT,,) was attributed to the optical properties and the material of the ATR configuration of the Micro CIRC LE® Open Boat Cell. It was postulated that the deeper penetration of the evanescent field into the sample at longer wave lengths (RTz) may have been responsible for this anomaly. It was also pointed out that the depth of penetration can be reduced by using alternate ATR configurations and ATR elements that have higher refractive indices such as AMTIR- 1 crystals. In the feasibility study presented in Chapter 1, the ATR element used was a cylindrical, ZnSe rod that has a low refractive index. In addition, the Micro CIRCLE® Open Boat Cell is a sampling device that may allow settling of slurries over the ATR rod leading to inconsistencies in accumulated data. In the studies undertaken subsequently and presented in Chapter 2 and here, the ATR element used was a conical, AMTIR-l ATR crystal that has a higher refractive index configured for in situ measurements. It is very likely that the combination of in situ measurements, the conical ATR configuration and the AMTIR- 1 ATR element with a higher refractive index 91 66- 05 DD % (w/w) Maleic Acid 01 q 'firv lammnlmlnllmnnnlnmn l annlnnmnlnnn 30 35 40 45 50 55 60 65 70 Temperature (°C) Figure 4.18 Solubility of maleic acid in water with respect to % maleic acid (w/w). 92 contributed to the consistency of the data extracted from the diverse set of IR parameters. Not all parameters considered were suitable for measurement of supersaturation as a result of some being independent of concentration. The existence of a clear distinction between parameters in terms of the dependence on concentration may be indicative of the unique structure of aqueous maleic acid. For example, consider PIR1 and PIIi‘2 that were extracted from the v(C=O) peak. PIRl is independent of concentration. Conversely, PIR2 is dependent on concentration. This may be indicative of the carbonyl bond being influenced by two types of H-bonds as illustrated in Figm'e 4.9. Evidence suggests that the folded structure of the maleic acid molecule in the crystalline state - due to intramolecular H-bonding - stays intact upon dissolution in concentrated solutions (0.3M) [5]. Consequently, the two carbonyl bonds of maleic acid in solution are most likely affected by both intramolecular and intermolecular H-bonding. Therefore, it can be postulated that the nature of the carbonyl bonds is manifested in the spectra as two mutually exclusive features, one independent (PEI) and the other dependent (PIRZ) on concentration. Again, such phenomena cannot be confirmed without other corroborative evidence by way of Raman and NMR spectroscopic data and experimental data using deuterated analogues of water and maleic acid. However, the broad scope of ATR FTIR spectroscopy for measurement of vital crystallization parameters and understanding crystallization systems is indisputable. 93 4.4 Conclusions The primary focus of this paper was to investigate the methods available to identify and isolate subtle features of IR spectra for utilization in measurement of crystallization parameters. Both derivative spectroscopy and deconvolution were successfully used to reveal features otherwise obscured by spectral artifacts. These features were primarily useful for quantitative measurements. The precision of the solubility measurements extracted from numerous IR parameters established the consistency of IR spectroscopy toward measurement of crystallization parameters. It is our firm belief that the techniques discussed herein can be used to help enhance features of many organic or inorganic crystallization systems toward identifying a parameter suitable for measurement of solubility and supersaturation. In addition to the versatility of ATR FTIR spectroscopy as a measurement tool, the spectra can be abundant with information that would provide some insight to the chemical natm'e of various systems. 4.5 References 1. Avram, M., Mateescu, GH. D., “Infrared Spectroscopy'Applications in Organic Chemistry”, Wiley-Interscience, New York, New York, (1966), Ch.7. 2. Hadzi, D., Bratos, S., in: “The Hydrogen Bond”, Eds. Schuster, P., Zundel, G., Sandorfy, C., North-Holland, New York, New York, (1976), Ch.12. 3. Davies, M., Sutherland, G. B. B. M., J. Chem Phys, 6, (1938), 755. 4. Hadzi, D., Chimia, 26, (1972), 7. 17. 18. 19. 20. 21. 22. 94 Hadzi, D., Novak, A, Spectrochim. Acta., 18, (1962), 1059. Lascombe, J ., Haurie, M., Josien, M. L., J. Chim. Phys, (1962), 1233. Flett, M. St. 0., Trans. Faraday Soc., 44, (1948), 767. Weltner Jr., W., J. Am. Chem. Soc, 77, (1955), 3941. Flett, M. St. C., J. Chem. Soc, (1951), 962. Kishida, S., Nakamoto, K., J. Chem. Phys, 41:6, (1964), 1558. Fukushima, K., Zwolinski, J ., J. Chem. Phys, 50:2, (1969), 737. Millikan, R. C., Pitzer, K S., J. Am. Chem. Soc, 80, (1958), 3515. Davies, M., Evans, J. C, J. Chem Phys, 20, (1952), 342. Grove, J. F., Williams, H. A., J. Chem. Soc, (1951), 877. Hadzi, D., Sheppard, N., Proc. Roy. Soc, London , A216, (1953), 247. Dunuwila, D. D., Berglund, K. A., Revised for publication in Ind. Eng. Chem. Fundamentals, (1995). Perkin Elmer IR Data Manager, User’s Manual, (1991). Dunuwila, D. D., Carroll II, L. B., Berglund, K. A., J. Cryst. Growth, 137, (1994), 561. Badger, R. M., Bauer, S. H., J. Chem. Phys, 5:11, (1937), 839. James, M. N. G., Williams, G. J. B.,Acta. Cryst, 330, (1974), 1249. The Aldrich Library of FP-IR Spectra, Vol. 1, Aldrich Chemical Company, Inc., Milwaukee, USA Bratoz, S., Hadzi, D., Sheppard, N ., Spectrochim. Acta., 8, (1956), 249. CHAPTER5 CONTINUING INVESTIGATIONS 5.1 Background 5.1.1 Control Strategy Definition of Supersaturation: It has been observed that maintenance of constant supersaturation during a crystallization process leads to improved crystal size distributions [1]. Supersaturation is defined in terms of the fundamental properties of solutions as given in Equation 2.6 of Chapter 2. However, for all practical purposes, supersaturation is approximated by the concentration difference or ratio with respect to the solubility with the underlying assumption of ideality [2]. This assumption is valid for sparingly soluble systems. For highly soluble systems, the approximation can lead to substantial misrepresentation of supersaturation due to increased nonideality of the solution. Generally, nonideal behavior of concentrated solutions, reflected by the activity coeficient (7), is due to extensive intermolecular and intramolecular interactions between the solute and the solvent. These interactions in turn configure the solution structure of the system. Then, it follows that the activity coeflicient is reflective of the solution structure at a molecular level. Therefore, a method for determining supersaturation that is reflective of the solution structure would be inherently more accurate. Conventional techniques such as refractometry, interferometry, viscometry 95 96 and densitometry, in addition to being unsuitable for in situ measurements, can measure only their respective bulk properties which are not sensitive to solution structure at a molecular level. In addition, conventional methods are not sensitive enough and their use can lead to gross misrepresentation of supersaturation. Conversely, infrared spectroscopy, which probes the vibrational structure of molecules is reflective of the solution structure and therefore is a viable technique to measure supersaturation reflective of solution activity. In Chapter 4, parameters sensitive to solution structure and supersaturation were distinguished from those that were sensitive to solution structure and not supersaturation, highlighting the specificity of IR spectroscopy. Further, the sensitivity of IR parameters (reflected by the signal gain) over a broad operating region was demonstrated in Chapter 4. Therefore, conceivable definitions for supersaturation include; Adam) = Pm -Pm,”l and mm) = Pm (5‘1) Pm,sol where Pm is any parameter discussed in Chapters 2, 3 and 4 in the context of measurement of supersaturation and solubility and Pm”, is the solubility with respect to P13. Since P3; is a parameter that reflects the fundamental structure of the solution, A(Pm) and 8(Pnz) will most likely contribute to a more accurate measurement of supersaturation. The sensitivity and the potential accuracy provided by ATM and 5(Pm) is anticipated to contribute toward the performance of a control scheme. It is also a direct measurement of supersaturation that does not require calibrations and therefore more practical. In efi'ect, a set point for a control scheme designed 97 to obtain a crystal size distribution (CSD) with a large average size can be envisioned as; A(Pm) = constant, or (5.2) 5(Pm) = constant which is illustrated in Figure 5.1. One of the primary objectives in controlling a crystallization process is to obtain a specified CSD. Excessive nucleation occurs when the system approaches its metastable linrit [3] resulting in CSDs with smaller average sizes. Therefore, if a larger average size is desired, it is imperative that the system is maintained well within the metastable zone; i.e., away from the limit. However, it is counterproductive to operate very close to the solubility line since this does not provide suficient driving force for growth resulting in unacceptably long batch times. Closed Loop Batch Crystallizer Control: A simple schematic of a batch crystallizer and the primary components of the envisioned control system is shown in Figure 5.2. The system consists of a jacketed crystallizer where a coolant fluid enters its jacket at temperature T,- and exits at T0. The flow rate is denoted by w. Supersaturation in the crystallizer measured by in situ ATR FTIR spectroscopy - A(Pm) or 5(Pm) - is measured and transmitted to the primary controller, E, which in turn adjusts the heat removal system, B, accordingly. In order to compensate for fluctuations in the heat removal system, a secondary controller, D, is suggested [4]. In this configuration, the heat flux, H , from the crystallizer measured by T,, T, and w is transmitted to the secondary controller which compensates for the fluctuations. Without a secondary controller, any fluctuation in the heat removal system 98 ) > Supersaturation > ————_———_——1 A(Pm) 0" 6(Pm) o Supersaturation, A(Pm) ) (mam ‘uonemresredns .9 t" o Temperature Time > 9 Figure 5.1 Control policy projected to maximize the mean sizeof a crystal size distribution from a cooling batch crystallization process. The policy is based on maintaininga constant supersaturation defined by the IR parameters given by A(Pm) or 5(Pm). A(Pm) = Pm - Pat”; and P 6(P )= m IR P H3301 where Pm is any parameter discussed in Chapters 2, 3and 4 in the context of measurement of supersaturation and solubility and Pm”, is the solubility with respect to Pm. * A(Pm) /. I [_________ Figure 5.2 A simple schematic of a batch crystallizer and the primary components of the envisioned control system. T, = Inlet coolant temperature of the jacketed crystallizer, To = Outlet coolant temperature, w = Coolant flow rate, MPH.) = Supersaturation, E = Primary controller B = Heat removal system, D = Secondary controller H = Heat flux from the crystallizer measured by T,, To and w. 100 can lead to a prolonged excursion in supersaturation from its setpoint. The reason for this is that the primary controller,without the aid of a secondary controller, does not take any corrective action until the distrn'bance is transmitted through the system consisting of several resistances. In the secondary control loop, heat flux fi'om the crystallizer is used as the measured variable as opposed to measuring the temperatm'e in the crystallizer since heat flux control has been proven to provide more stable control systems [5]. The block diagram of the cascade control system is given in Figure 5.3 (a). Although, a good heat removal system is tantamount to the stable operation of the overall control scheme, its design is not expected to be a limiting factor considering the resources available for its design [6]. Therefore, the heat removal system is not specified. The task more substantial in terms of relevancy and more challengingwillbe the design of the primary loop controlling supersatm'ation. In order to consummate the goal of designinga stable supersaturation control scheme (primary loop), it is not necessary to know the dynamics of the secondary loop, since the two loops can be analyzed independent of each other. The standard procedure in designing cascade control systems is to first establish the control dynamics for a stable secondary loop. Subsequently, the secondary loop can be normalized to determine the stability of the primary loop. This is demonstrated between Figures 5.3 (a) and 5.3 (b). However, in this case its reasonable to consider normalizingthe secondary loop for heat flux control without specifying the heat removal system taking comfort in the resources available for its design in terms of hardware, literature and specialists. 101 .2: 5 peaches—a mm 98— 26583 3588 use: 3238.25 23. .8893 35:8 £533 55 mo 38521 Joe—m rem warm snag _ , 15.85294 _ 3V . n84 888m .5380 T lfi H 3.530 $8.8m 395 no: 853mm , H 358$” a 8m _ , .532 - 555m seas 585550 853.555 .558 gees: _ - sees 525% _ a§82_ - . N37m ”:33 h _ Adv 388m 55580 9 .5380 3683* So: 588m basin 8.323 I ewe mam awefififieusw L 5.58 102 Thus, the hitherto unventmed, yet promising aspect of crystallizer design and control, which is design and control based on in situ measurement of supersaturation, can be promptly attended to as it is more relevant to and more substantial in the field. One concern in using in situ ATR FTIR spectroscopy for control of batch crystallization processes may be whether the frequency of measurement would be sufficient for implementation in control of a highly labile process such as batch crystallization. The scan rate of the supersaturation profiles presented in Chapter 3 (Figures 3.5 and 3.7) was 0.05 sec". In fact, the desupersaturation of aqueous citric acid, which is one of the fastest growing systems known, was measured with remarkable point to point resolution and is shown in Figure 5.4. Desupersaturation of aqueous citric acid was measured using a protocol similar to that used with maleic acid at a scan rate of 0.008 sec". The citric acid solution was not seeded. In view of a control scheme with high performance, a scan rate of 0.1 sec'1 is realizable. Clearly, the data suggest that the frequency of measurement will not be a limiting factor in the development of a control scheme. Crystallization Modeh The particle number-size distribution theory introduced by Randolph and Larson [7] and the fundamental supersaturation balance discussed by Mullin and Nyvlt [8, 9, 10] provide powerful models to evaluate crystallization processes. The model derived from the particle number-size distribution theory is based on the population balance (number of particles per unit volume per length range) and takes nucleation and growth of crystals into account. The fundamental supersaturation balance is derived taking into consideration that the driving 103 3'1'." O TIRprofile + Solubility TIR 144 m1 1 I a] nnnnnn Ill nnnnn IAAAJIAJLLILJJJ 5 10 15 20 25 30 35 40 45 Temperature (°C) Figure 5.4 A supersaturation profile of aqueous citric acid represented in terms of TIR. transmittance at 3284 cm'1 TIR = , 1 transmittance at 1190 cm' 104 force for all crystallization phenomena such as nucleation and growth is supersaturation. Thus, kinetic parameters for nucleation and growth are nested in these models giving them both predictive and descriptive capabilities. Another valuable set of equations commonly used in crystallization analysis is the moments of the particle number-size distribution [7]. The moments have been utilized to circumvent a basic dimensional incompatibility between the population balance and the transport equations for mass, momentum and energy. The particle phase space consists of a set of internal coordinates in adchtion to the three spatial dimensions. This incompatibility complicates the simultaneous solution of the population balance with the transport equations. The most significant internal coordinate in the particle phase space is the particle size. Therefore, the population density distribution is integrated over the particle size internal coordinate reducing the dimensionality of the population balance to that of the transport equations. The moments that give the total particle properties (presented shortly) upon integrating the population balance over the particle size coordinate are given by; m ,0) = Tn(L,t)LjdL, j = 0, 1,2, (5.3) 0 where n is the population density distribution and L is the particle size. The application of these models in engineering simulations and process analysis and control has been limited by the unavailability of a reliable and sensitive technique to obtain kinetic data in situ. In crystallization processes, in situ measurement is extremely important since 105 the hydrodynamic conditions set by the crystallizer dimensions, flow patterns induced by the propeller and the suspended crystal slurry can significantly alter kinetic parameters from that obtained under controlled laboratory experiments [1 1]. Theoretically, kinetic parameters can be extracted from the desupersaturation rate of the solution monitored by in situ ATR FTIR spectroscopy. An attempt was made to extract the crystallization kinetic parameters of aqueous maleic acid by least squares minimization between the measured supersaturation profile and the supersaturation profile calculated using the supersaturation balance which has a structure similar to the moment equations. Crystallization is a particulate process where the number of particles generated over the batch time plays a significant role in affecting the course of the process. Therefore, it was extremely dificult to obtain a consistent set of crystallization kinetic parameters using supersaturation alone since it does not provide number related information. It is anticipated that in situ measurement of supersatm'ation using ATR FTIR spectroscopy and measurement of the transient CSD using a light scattering device in combination will be a more promising approach for the estimation of kinetic parameters. The model most convenient for the extraction of kinetic parameters using supersaturation measurements and CSD measurements is the set of moment equations [7]. The population balance for a well mixed, constant volume batch crystallizer is given by, 913(6):) _ 3: 6L ’ O (5.4) 106 where G is the linear growth rate of crystals. In order to obtain the moments of the distribution, the population balance is integrated over the particle size coordinate as follows; T[%”+ 32%") = 0]LJ'dL, j = 0,1, 2, (5.5) 0 The resulting set of moments are given by; iti—OJB" = JGmJ-_1, j =0,],2,... (5.6) where 0’ = 0 for j at O. The time dependent moments are as follows; =__dNn = 30 m dt dm '3?" d3. : lg : mlG (57) dt is; dt dm. '21? where the number rate of nucleation, B”, the number of seeds, N,, and the linear crystal growth rate, G, are defined in terms of the frequently used power model as; 3°: kuplL ———[k,,,pA(Pm)" wk, ,,-M’"A(PIR)”] c n,0 kc, 3495‘ M__,__o% N3: . 10.0.12 where N, is the total number of nuclei, L is the total length of crystals, A is thetotalareaofcrystaIs,Misthetotalmassofcrystalsinsuspension, k, is G: k ,A(Pm)g (5.8) 107 the area shape factor, k, is the volume shape factor, p, is the crystal density, Lmo is the size of nuclei at birth, Lw is the initial size of seed crystals, W“, is the initial mass of seed crystals, k”, is the primary nucleation rate constant, k M is the secondary nucleation rate constant, k g is the growth rate constant, 11,, is the order of primary nucleation, n, is the order of secondary nucleation and g is the order of growth. In order to solve the moment equations, the operating conditions should be defined and the mass balance should be specified. The mass balance is specified as; “(1332) = _ dPIR,sol _ M 5.9 dt dt dt ( ) The time dependent population density distribution can be recovered by the following approximation [7]; n(L) = A'lm (5.10) where m is the moment vector and the elements of the matrix A is given by; a j, = LiALk (5.11) where L, is the midpoint of a size range AL, It has been shown that four to five moments satisfactorily reproduce an original distribution [7]. The higher moments unduly weight the distributions toward the larger sizes. Control System Equations; State Variable Form: State variable form is particularly convenient for the description of complicated systems for which a standard, compact notation is especially helpful [12]. The state representation of systems, that uses integrators, is useful not only in control system analysis, but for simulation and analysis of systems of all kinds. A state variable description of an nth-order system, in the Laplace domain, 108 involves n integrators and the corresponding outputs which are the state variables. The inputs of each of the integrators is given as a linear combination of the state signals, X", and the system inputs, 12,-; “1(8) = 011X1(S) + 012X2(8) + ... + (1th (8) + b11R1(S) + ... + “1.121(8) 8X2(S) = a21X1(S) + 022X2(3) + + 02an (S) + b21R1(8)+ + bZiRl(s) (5 12) Ex" (3) = an1X1(S) + an2X2(8) + + 0,,an (3) + bn1R1(S) + ... ‘1" built! (8) In the time domain these are a set of n firstorder linear differential equations; “(l—t1 = (11111 + a12x2 + ...+ ahxn + bur, + +b1lrl- dxz dt = amxl +a22x2 + ...+a2,,x,, +521r, +...+ 52,1;- (5.13) -d—tn = anlxl +an2x2 + ...+a,,,,x,, + bnlr, + ...+b,,,-r,- In matrix notation the state variable can be compiled as; "’51- “‘1- F011 012 aln-le- P511 blinrl- ixz=izzazl 022 021 x2+b21 “bzirz dt 5 E 1 3 '- i i E ' i E _xnj _in_ _anl an2 ann__xnd _bnl bni‘_ri_, or (5.14) dx —=i=Ax+Br dt The system outputs are related to the state variables through the output equations; 3’1 011 012 Cb: x1 ym cm] cm2- or y :0: cmn xn (5.15) 109 The time course of the state variables are described by the state equations. This is a standard form of representing any system described by linear constant-coefficient integrodifl‘erential equations. Often, systems consisting of a mixed collection of simultaneous system equations, some of first order, some of second order, some involving running integrals, etc., are manipulated to place them in this standard form. The primary advantage is that systematic methods of analysis such as frequency response analysis can be implemented on involved problems quite efficiently. For example, it may be desirable to determine the contributions of each of the state variables to the overall process gain of the system. In this exercise, the more significant state variables can be isolated from those that are much less significant Thus, more insignificant state variable can either be neglected or considered as a disturbance input to the system. The latter scenario is described by the following equation assuming that the contribution of the state variable x, is negligible. {-551- Parr “13 arnvxrl [bu bit-7'11 ’0127 x3 031 033 031 J"3 bar b3i "3 + “223:2 (516) in, _anl an3 annJ_xn_ Lbnl bnid_ri_ _an24 This analysis gives process development engineers an insight to the process parameters most important toward design and development. In crystallization the process state variables include, supersaturation, CSD, slurry density, total crystal surface area, nucleation rate, growth rate and heat flux State variable analysis briefly outlined above will help identify the relative importance of process parameters related to each of the state variables at various operating conditions. The 110 relative merits of in Situ measurement of supersaturation and in situ measurement of CSD can be analyzed. The set of equations that describe a well mixed constant volume batch crystallizer is given by Equations 5.7, 5.8, 5.9 and 5.10. The differential equations are highly nonlinear. However, a methodology for the systematic characterization of dynamic systems by state variable analysis has been developed only for linear systems as presented above. Thus, the convenience of state variable analysis is significant motivation for approximating nonlinear systems with linear ones. One technique commonly used is linearization around a judiciously chosen operating point using a Taylor series expansion [4]. The success of the linear approximation depends on the deviation of the system fi‘om the chosen operating point. The smaller the deviation the better the approximation. Therefore, process simulations will have to be used in aid of choosing the most appropriate operating point. In the rare event that linearization around an operating point fails to provide a reasonable approximation, more dificult methods of nonlinear analysis may have to be used [4]. 5.1.2 Solution Structure In Chapter 4 a reasonable amount of evidence was presented in the context of inferring the solution structure of aqueous maleic acid in supersaturated solutions. The presented evidence suggest that the folded structure of crystalline maleic acid remains intact in highly concentrated solutions. However, a concrete analysis on the nature of the solution state structure of aqueous maleic acid in supersaturated solutions was beyond the scope of the investigation discussed in Chapter 4. 111 The depth of the investigation can be extended using both Raman and NMR spectroscopy in addition to FTIR spectroscopy. The difi‘erent techniques can reveal complementary information. Further, utilization of the deutero analogues of both water and maleic acid in experiments similar to those discussed in Chapter 4 can provide a host of confirmatory information [13, 14, 15]. The deuteration of a particular bond results in a significant red shift in its vibrational frequencies due to the heavy atom effect. For example, the v(O-D) vibration of the deuterated acetic acid is observed at 2299 cm‘1 in comparison to the v(O-H) vibration of acetic acid that occurs at 3140 cm’1 [13]. The 6(O-D) vibration is observed at 1046 cm’1 compared to the 5(O-H) vibration at 1294 cm". Consequently, experimentation with selectively deuterated maleic acid and deuterated water in judiciously chosen combinations can help resolve many of the ambiguities of the analysis provided in Chapter 4. For example, use of maleic acid in heavy water will clear the v(O-H) stretching region of any contribution hour the v(O-H) vibration of water thus making it much easier to identify dynamics of the v(O-H) vibration of maleic acid. In Chapter 4, the peak near 1629 cm'1 was assigned to the v(C=C) stretching mode. However, the strength of the peak is not consistent with the vibrational mode. The strength of the IR active vibrational modes depend on the strength of the permanent dipole moment of the mode. Eventhough the maleic acid molecule is not perfectly symmetric around the 112 (C=C) bond, its relative symmetry prevents extensive polarization of the (C=C) bond. Therefore, the peak intensity of the v(C=C) mode is expected to be week. In fact, the v(C=C) mode is nonexistent in the IR spectrum of fumaric acid, which is the trans configuration of maleic acid, since it has a center of symmetry around the (C=C) bond. Consequently, the most likely event is that the strong band around 1629 cm‘1 is constituted by contributions from the v(C=C) mode and other modes of maleic acid. The other contributors could be overtones of lower fi'equency modes of maleic acid. The origins of these overtone bands can be scrutinized by experimentation with selectively deuterated maleic acid and deuterated water in judiciously chosen combinations. Separation of this convoluted band using deuterated species can provide additional information toward deducing the solution structure of aqueous maleic acid. Ambiguities resulting from band convolution of other identified modes can be resolved in similar fashion. In addition, assignments can be made on unidentified band and shoulders. Studies with complementary spectroscopic techniques such as Raman and NMR can be conducted with deuterated and non-deuterated species adding to the wealth of information available to study the solution structure. 5.1.3 Effect of impurities In industrial crystallization, solutions invariably contain impurities within detectable limits. In some crystallization schemes, impurities may also be added to instigate desirable conditions. Impurities can significantly alter the solubility characteristics and crystallization kinetics of the 113 primary solute. In the presence of an impmity one of three situations is likely to arise. The impmity may react with the primary solute and completely alter the nature of the system, both chemically and physically. The presence of an impurity may drive the primary solute into supersaturation and thereby cause the primary solute to precipitate out of solution. Lastly, the presence of an impurity may cause the system to be undersaturated with respect to the primary solute. In the literature, the latter two effects are identified as 'salting—out' and 'salting-in', respectively. In addition, kinetics can be afl‘ected by the adsorption of impurity molecules on growing crystal surfaces. In industrial crystallization processes, monitoring of impmity (either contained or added) levels and/or their effects, is crucial for a successful operation. ATR FTIR spectroscopy, in addition to having the capability to measure solubility and supersaturation, can be used to monitor the presence of additives or impurities and consequent reactions (reactive crystallization schemes included). Infrared spectra, that reflect the vibrational and rotational structure of the analytes, can provide more than one piece of information unlike traditional solubility and supersaturation measurement techniques that measure a single parameter calibrated for the measurement of solubility and supersaturation. The advantage is that necessary information can be obtained simultaneously using the same technique, in situ, without resorting to multiple techniques that are variable specific. The potential for the application of in situ ATR FTIR spectroscopy to study impurity efl‘ects remain to be explored. 114 5.2 Conclusions The utility of in situ ATR FTIR spectroscopy for the measurement of important crystallization parameters such as supersaturation, solubility and metastable limit was demonstrated in the previous chapters. The feasibility of the technique toward process development and eventually control of batch crystallization processes that will provide for improved product quality/reproducibility, time savings and cost containment was established. However, its potential for implementation in control of crystallization processes, although promising, remains to be proven. In chapter 4 substantial evidence demonstrating the potential of in situ ATR FTIR spectroscopy towards deducing solution structure in supersaturated solutions was presented. Utilization of in situ ATR FTIR spectroscopy in studying the effects of impurities on crystallization systems is yet another avenue remaining to be explored. Therefore, it is recommended that the implementation of control using in situ measurement of supersaturation using ATR FTIR spectroscopy, utilization of IR spectroscopy for understanding solution structure in supersaturated solutions and investigation of efl‘ects of impurities on crystallizing systems using ATR FTIR spectroscopy be pursued in the futm'e. 5.3 References 1. Jones, A G., Mullin, J. W., Chem. Eng. Sci, 29, (1974), 105. . Garside, J ., Chem. Eng. Sci, 40:1, (1985), 3. 3. Mullin, J. W., “Crystallization”, Butterworth-Heineman, Oxford, England, (1993). 10. 11. 12. 13. 14. 15. 115 Coughanowr, D. R., “Process Systems Analysis and Control”, McGraw-Hill, Inc., New York, USA Radcliffe, 0., Professor, Mechanical Engineering, Michigan State University. Perry, R. , Chilton, Perry’s Handbook of Chemical Engineering, McGraw Hill, New York, 6th Edition. Randolph, A. D., Larson, M. A., “Theory of Particulate Processes”, Academic Press, Inc., New York, USA, (1988), ch. 3. Mullin, J. W., Nyvlt, J ., Chem. Eng. Sci, 26, (1971), 369. Nyvlt, J ., Colin. Czech. Chem. Commun., 30, (1965), 2269. Nyvlt, J ., Mullin, J. W., Chem. Eng. Sci, 25, (1970), 131. N. S. Tavare, Separation and Purification Methods, 22:2, 93, (1993). Hostetter, G. H., Savant, Jr., C. J ., Stefani, R. T., “Design of Feedback Control Systems”, Holt, Rinehart and Winston, New York, USA, (1982), ch. 7. Kishida, S., Nakamoto, K, J. Chem. Phys, 41:6, (1964), 1558. Fukushima, K, Zwolinski, J ., J. Chem. Phys, 50:2, (1969), 737. Bratoz, S., Hadzi, D., Sheppard, N., Spectrochim. Acta., 8, (1956), 249. APPENDIX APPENDIX Table A1 Data for Figure 2.4 2.27 0.025 2.27 0.036 4.37 0.024 4.37 0.037 13.31 0.035 13.31 0.053 22.16 0.043 15.02 0.053 23.54 0.040 22.16 0.075 26.95 0.084 23.54 0.078 29.35 0.068 29.35 0.104 30.48 0.077 30.48 0.104 33.12 0.107 34.42 0.155 35.75 0.105 41.77 0.201 35.75 0.100 44.84 0.233 35.75 0.107 44.84 0.245 38.51 0.135 47.65 0.246 40.79 0.127 48.67 0.281 40.79 0.148 50.18 0.285 40.79 0.126 50.18 0.297 43.07 0.165 51.85 0.305 44.87 0.156 52.36 0.288 44.87 0.157 54.51 0.338 44.87 0.173 54.51 0.327 47.07 0.199 54.72 0.355 48.66 0.200 56.45 0.343 48.66 0.196 57.24 0.386 48.66 0.188 58.21 0.388 50.48 0.230 58.21 0.384 51.87 0.221 59.50 0.413 51.87 0.220 59.79 0.392 51.87 0.239 61.30 0.421 53.48 0.258 61.30 0.431 54.65 0.261 61.54 0.443 54.65 0.257 62.65 0.423 56.16 0.287 63.43 0.475 57.18 0.307 63.98 0.467 57.18 0.280 63.98 0.468 59.51 0.319 65.07 0.500 59.51 0.348 65.20 0.458 61.53 0.370 66.28 0.493 61.53 0.348 66.29 0.508 68.36 0 535 116 117 Table A2 Data for Figures 2.5 and 2.6 Temperature (°C) RT. at 10 °c Temperature (°C) RT. at 30 °C 2.27 0.043 2.27 0.056 4.37 0.051 4.37 0.064 13.31 0.089 13.31 0.117 22.16 0.160 15.02 0.127 23.54 0.113 22.16 0.239 26.95 0.277 23.54 0.289 29.35 0.328 29.35 0.405 30.48 0.403 30.48 0.472 33.12 0.436 34.42 0.641 35.75 0.547 41.77 1.069 35.75 0.528 44.84 1.483 35.75 0.551 44.84 1.301 38.51 0.675 47.65 1.632 40.79 0.888 48.67 1.719 40.79 0.927 50.18 2.165 40.79 0.853 50.18 1.900 43.07 0.983 51.85 2.006 44.87 1.284 52.36 2.316 44.87 1.265 54.51 2.501 44.87 1.289 54.51 2.974 47.07 1.380 54.72 2.559 48.66 1.768 56.45 3.047 48.66 1.847 57.24 3.043 48.66 1.803 58.21 3.336 50.48 1.821 58.21 3.946 51.87 2.405 59.50 3.568 51.87 2.333 59.79 3.826 51.87 2.390 61.30 4.642 53.48 2.321 61.30 4.091 54.65 2.965 61.54 4.255 54.65 3.031 62.65 4.527 56.16 2.797 63.43 4.791 57.18 3.915 63.98 5.611 57.18 3.771 63.98 4.792 59.51 4.871 65.07 5.326 59.51 4.801 65.20 5.454 61 .53 5.512 66.28 6.482 61.53 5.589 66.29 5.451 68.36 6.491 118 Table A3 Data for Figures 2.7 and 2.8 Temperature (°C) RT, etdev In RT. RT; 10 0.266 0.007 3.410 15 0.319 0.009 4.317 20 0.382 0.009 5.305 25 0.429 0.008 6.735 30 0.484 0.009 8.271 35 0.532 0.012 9.597 etdev in RT; 0.143 0.215 0.328 0.454 0.739 1.074 119 Table A4 Data for Figures 3.3 and 3.5 TR TR Supersaturation Temperature TR Temperature Supersaturation) Temperature profile (°C) Solubility (°C) profile (°C) continued 52.70 4.821 66.50 7.508 62.21 7.296 56.60 5.611 66.44 7.425 62.05 7.378 60.51 6.443 66.38 7.529 61.90 7.252 64.51 7.408 66.32 7.611 61.74 7.345 66.47 7.874 66.24 7.501 61.59 7.332 TR 66.16 7.459 61.44 7.239 Temperature Metastable 66.08 7.531 61.29 7.354 ° Limit 65.98 7.495 61.15 7.357 60.13 7.28 65.89 7.515 61.00 7.269 58.26 6.72 65.78 7.352 60.85 7.230 54.59 5.92 65.68 7.533 60.70 7.210 53.68 5.51 65.57 7.471 60.55 7.250 50.67 5.03 65.45 7.452 60.40 7.180 65.33 7.537 60.30 7.240 65.20 7.449 60.15 7.210 65.07 7.441 60.00 7.270 64.94 7.414 59.75 7.240 64.81 7.495 59.60 7.200 64.67 7.353 59.45 7.250 64.53 7.349 59.34 7.202 64.38 7.394 59.69 7.326 64.23 7.461 60.32 6.976 64.09 7.475 60.60 6.960 63.93 7.508 60.64 6.997 63.78 7.280 60.64 6.774 63.63 7.504 60.64 6.771 63.47 7.378 60.57 6.836 63.32 7.474 60.57 6.716 63.16 7.275 60.55 6.676 63.00 7.316 60.27 6.605 62.84 7.425 60.23 6.582 62.68 7.273 60.35 6.621 62.53 7.440 59.94 7.130 62.37 7.345 60.15 7.100 120 Table A5 Data for Figure 3.6 Temperature Temperature (°C) Temperature Temperature (°C) Parabolic (°C) (°C) Time Parabolic Time cooling Time Int. parabolic Time Linear (sec) cooling (sec) continued (sec) cooling (sec) cooling 0 63.04 1400 58.69 0 63.56 0 62.68 40 63.04 1440 58.42 40 63.45 20 62.26 80 63.03 1480 58.15 80 63.34 40 61.83 120 63.03 1520 57.86 120 63.21 60 61.40 160 63.02 1560 57.56 160 63.08 80 60.97 200 63.00 1600 57.26 200 62.94 100 60.54 240 62.98 1640 56.95 240 62.79 120 60.11 280 62.94 1680 56.63 280 62.64 140 59.68 320 62.90 1720 56.30 320 62.47 160 59.25 360 62.85 360 62.30 180 58.81 400 62.79 400 62.12 200 58.38 440 62.73 440 61.93 220 57.95 480 62.65 480 61.73 240 57.51 520 62.57 520 61 .53 260 57.08 560 62.48 560 61.31 280 56.64 600 62.38 600 61.09 300 56.21 640 62.27 640 60.86 680 62.16 680 60.62 720 62.04 720 60.38 760 61.90 760 60.12 800 61.76 800 59.86 840 61.62 840 59.59 880 61.46 880 59.31 920 61.30 920 59.02 960 61.13 960 58.72 1000 60.95 1000 58.42 1040 60.76 1040 58.10 1080 60.56 1080 57.78 1120 60.36 1120 57.45 1160 60.14 1160 57.11 1200 59.92 1200 56.77 1240 59.69 1240 56.41 1280 59.46 1280 56.05 1320 59.21 1320 55.68 1360 58.96 1360 55.30 121 Table A6 Data for Figure 3.7 TR TR Supersaturation Temperature Supersaturation Temperature Profile Profile (°C) continued 56.13 5.526 61.54 6.988 56.30 5.609 61.62 6.954 56.47 5.667 61.69 6.931 56.63 5.716 61.76 6.837 56.79 5.714 61.84 6.922 56.95 5.730 61.90 6.970 57.11 5.809 61.97 6.854 57.26 5.758 62.04 6.960 57.41 5.823 62.10 7.048 57.56 5.826 62.16 6.944 57.71 5.918 62.22 7.070 57.86 5.926 62.27 7.065 58.00 5.926 62.33 7.070 58.15 6.033 62.38 7.089 58.29 6.039 62.43 7.079 58.42 6.021 62.48 7.099 58.56 6.079 62.52 7.201 58.69 6.190 62.57 7.036 58.83 6.195 62.61 7.191 58.96 6.212 62.65 7.159 59.09 6.270 62.69 7.137 59.21 6.255 62.73 7.261 59.33 6.309 62.76 7.203 59.46 6.312 62.79 7.160 59.58 6.386 62.82 7.210 59.69 6.347 62.85 7.226 59.81 6.501 62.88 7.307 59.92 6.517 62.90 7.287 60.03 6.486 62.92 7.312 60.14 6.517 62.94 7.286 60.25 6.522 62.96 7.114 60.36 6.447 62.98 7.194 60.46 6.482 62.99 7.366 60.56 6.707 63.00 7.278 60.66 6.599 63.01 7.340 60.76 6.619 63.02 7.215 60.85 6.729 63.02 7.176 60.95 6.770 63.03 7.316 61.04 6.753 63.03 7.282 61.13 6.858 63.03 7.294 61.21 6.757 63.03 7.289 61.30 6.883 63.04 7.322 61.38 6.765 63.04 7.270 61.46 6.910 63.04 7.228 122 Table A7 Data for Figure 3.8 Tn Tn Supersaturation Temperature Supersaturation Temperature Profile (°C) Profile (°C) continued 55.296 5.3865 60.744 6.6312 55.487 5.4195 60.862 6.7642 55.677 5.5388 60.978 6.8338 55.864 5.4821 61.092 6.7956 56.049 5.5299 61.204 6.8988 56.232 5.6675 61.314 6.8647 56.412 5.658 61.422 6.8572 56.591 5.7099 61.528 6.8928 56.767 5.7356 61.631 6.7263 56.942 5.7091 61.733 6.946 57.114 5.7315 61.833 7.1089 57.285 5.801 61.93 7.1234 57.453 5.8375 62.026 7.1212 57.619 5.8934 62.119 7.1956 57.783 5.9689 62.211 7.15 57.945 5.8844 62.3 7.196 58.104 6.0924 62.387 7.1268 58.262 6.0357 62.472 7.1791 58.418 5.9942 62.556 7.1612 58.571 6.1303 62.637 7.0954 58.722 6.0929 62.716 7.1826 58.872 6.1093 62.793 7.2122 59.019 6.1969 62.868 7.2826 59.164 6.3091 62.941 7.3143 59.307 6.3879 63.011 7.1401 59.448 6.336 63.08 7.1582 59.587 6.262 63.147 7.1531 59.724 6.3609 63.212 7.1343 59.859 6.4884 63.274 7.1327 59.991 6.4618 63.335 7.1691 60.122 6.4843 63.394 7.0479 60.25 6.4718 63.45 7.235 60.377 6.58 63.505 7.1843 60.501 6.5267 63.557 7.2077 60.624 6.6256 Data for Figure 3 Table A8 Ta Temperature Supersaturation (°C) Profile 56.207 5.6364 56.643 5.7073 57.078 5.7531 57.513 5.7942 57.947 5.9737 58.38 6.1093 58.813 6.1557 59.246 6.4522 59.677 6.4576 60.108 6.7233 60.539 6.7861 60.969 7.0622 61.398 7.1281 61.827 7.0984 62.255 7.1642 62.682 7.2312 123 124 Table A9 Data for Figure 3.10 CSD (Stweight) CSD (Stweight) Intermediate CSD (Straight) Parabolic parabolic Linear Size cooling cooling cooling (microns) profile profile profile 180 0.00 0.52 0.44 265 0.40 1.40 2.08 373 4.10 13.16 45.84 463 4.90 9.58 15.88 550 9.20 14.80 14.78 655 13.10 13.62 5.28 780 23.70 21.84 1.90 925 18.00 12.88 0.00 1000 13.30 5.14 0.00 125 Table A10.1 Data for Figure 3.11 Tn TR Concentration Concentration Temperature line Temperature line (°C) 60 % wlw (°C) 60 % wlw 56.89 4.827 56.90 4.795 56.86 4.778 56.86 4.833 56.82 4.764 56.81 4.811 56.76 4.838 56.76 4.856 56.71 4.801 56.70 4.853 56.64 4.807 56.64 4.882 56.57 4.774 56.57 4.864 56.50 4.772 56.50 4.822 56.42 4.750 56.43 4.784 56.33 4.762 56.35 4.758 56.24 4.694 56.27 4.794 56.15 4.739 56.19 4.839 56.05 4.762 56.10 4.790 55.95 4.783 56.01 4.785 55.84 4.714 55.91 4.773 55.74 4.772 55.81 4.850 55.63 4.768 55.71 4.735 55.52 4.738 55.61 4.770 55.41 4.795 55.50 4.738 55.29 4.744 55.39 4.791 55.18 4.779 55.28 4.840 55.06 4.729 55.16 4.859 54.94 4.823 55.04 4.814 54.82 4.830 54.92 4.792 54.70 4.761 54.80 4.779 54.58 4.774 54.68 4.745 54.46 4.727 54.55 4.738 54.34 4.753 54.42 4.813 54.22 4.790 54.29 4.705 54.09 4.787 54.16 4.672 53.97 4.706 54.03 4.798 53.84 4.770 53.89 4.736 53.71 4.797 53.76 4.736 53.59 4.780 53.62 4.753 53.46 4.685 53.48 4.806 53.33 4.682 53.34 4.729 53.20 4.751 53.20 4.705 53.06 4.712 53.06 4.693 52.93 4.739 52.91 4.736 52.79 4.703 52.77 4.745 52.65 4.826 52.62 4.710 52.51 4.698 52.47 4.715 52.36 4.760 52.33 4.761 52.21 4.701 52.18 4.727 52.06 4.686 52.03 4.749 51.90 4.715 51.88 4.728 126 Table A10.2 Data for Figure 3.11 TR TR Concentration Concentration Temperature line Temperature line (°C) 61 7. WM (°C) 61 % wlw 58.52 5.200 58.54 5.266 58.52 5.228 58.53 5.387 58.51 5.283 58.51 5.270 58.48 5.250 58.49 5.167 58.45 5.221 58.46 5.184 58.42 5.252 58.42 5.303 58.37 5.314 58.38 5.258 58.32 5.224 58.33 5.224 58.26 5.299 58.28 5.221 58.20 5.207 58.23 5.307 58.13 5.196 58.17 5.246 58.05 5.238 58.10 5.251 57.97 5.247 58.03 5.217 57.89 5.313 57.96 5.198 57.80 5.291 57.88 5.245 57.70 5.288 57.80 5.234 57.60 5.184 57.71 5.202 57.50 5.169 57.62 5.191 57.39 5.221 57.53 5.260 57.28 5.153 57.43 5.248 57.16 5.188 57.33 5.094 57.04 5.271 57.23 5.239 56.92 5.243 57.12 5.255 56.80 5.142 57.01 5.168 56.67 5.137 56.89 5.278 56.55 5.131 56.78 5.228 56.41 5.180 56.66 5.211 56.28 5.149 56.53 5.183 56.15 5.226 56.41 5.161 56.01 5.211 56.28 5.224 55.87 5.148 56.15 5.242 55.73 5.180 56.01 5.201 55.59 5.181 55.88 5.156 55.45 5.197 55.74 5.180 55.31 5.163 55.60 5.177 55.16 5.163 55.46 5.194 55.02 5.195 55.32 5.191 54.87 5.182 55.17 5.152 54.73 5.140 55.03 5.104 54.58 5.075 54.88 5.173 54.43 5.182 54.73 5.127 54.28 5.181 54.58 5.203 54.13 5.106 54.43 5.219 53.99 5.121 54.28 5.116 53.84 5.127 54.13 5.200 53.69 5.064 53.97 5.168 127 Table A10.3 Data for Figure 3.11 Tn TR Concentration Concentration Temperature line Temperature line (°C) 62 % wlw (°C) 62 % wlw 60.34 5.685 60.29 5.832 60.33 5.703 60.32 5.768 60.32 5.730 60.33 5.704 60.30 5.754 60.33 5.786 60.27 5.710 60.32 5.760 60.24 5.777 60.30 5.756 60.20 5.795 60.27 5.719 60.15 5.655 60.24 5.681 60.10 5.751 60.19 5.709 60.04 5.716 60.14 5.690 59.97 5.759 60.09 5.681 59.90 5.670 60.02 5.710 59.83 5.637 59.95 5.712 59.75 5.622 59.87 5.751 59.66 5.696 59.79 5.661 59.57 5.712 59.70 5.772 59.47 5.656 59.61 5.614 59.37 5.573 59.51 5.759 59.27 5.730 59.40 5.734 59.16 5.707 59.30 5.734 59.04 5.700 59.18 5.725 58.93 5.691 59.07 5.634 58.81 5.699 58.94 5.682 58.68 5.765 58.82 5.594 58.55 5.666 58.69 5.639 58.42 5.675 58.56 5.720 58.29 5.670 58.43 5.661 58.16 5.613 58.29 5.677 58.02 5.654 58.15 5.692 57.88 5.596 58.01 5.640 57.74 5.651 57.87 5.608 57.59 5.554 57.72 5.655 57.45 5.625 57.58 5.719 57.30 5.568 57.43 5.721 57.15 5.673 57.28 5.644 57.00 5.640 57.13 5.584 56.85 5.538 56.98 5.529 56.70 5.506 56.83 5.577 56.55 5.634 56.67 5.613 56.39 5.577 56.52 5.551 56.24 5.626 56.36 5.581 56.09 5.561 56.21 5.536 55.94 5.598 56.05 5.632 55.79 5.565 55.90 5.541 55.63 5.565 55.74 5.612 55.48 5.498 55.59 5.489 128 Table A10.4 Data for Figure 3.11 TR TR Concentration Concentration Temperature line Temperature line (°C) 63 7. w/w (°C) 63 % wlw 62.20 6.059 62.22 6.138 62.21 6.269 62.23 6.211 62.21 6.159 62.22 6.247 62.21 6.195 62.21 6.191 62.19 6.068 62.19 6.287 62.16 6.180 62.16 6.141 62.13 6.108 62.12 6.134 62.08 6.216 62.07 6.084 62.03 6.205 62.01 6.153 61.97 6.141 61.95 6.141 61.91 6.173 61.88 6.181 61.83 6.134 61.80 6.193 61.75 6.123 61.72 6.063 61.66 6.139 61.63 6.190 61.57 6.145 61.53 6.122 61.47 6.060 61.43 6.106 61.37 6.139 61.33 6.087 61.26 6.045 61.22 6.231 61.15 6.146 61.10 6.163 61.03 6.265 60.99 6.153 60.91 6.130 60.86 6.089 60.78 6.144 60.74 6.216 60.65 6.117 60.61 6.107 60.51 6.075 60.48 6.252 60.38 6.143 60.35 6.148 60.24 6.114 60.21 6.164 60.09 6.158 60.07 6.185 59.95 6.198 59.94 6.098 59.80 6.082 59.79 6.173 59.65 6.049 59.65 6.095 59.49 6.056 59.51 6.063 59.34 6.100 59.36 6.028 59.18 6.050 59.22 6.126 59.03 6.014 59.07 6.041 58.87 6.140 58.92 6.106 58.71 6.091 58.77 6.038 58.55 6.038 58.62 6.002 58.38 6.026 58.47 5.990 58.22 6.062 58.32 6.036 58.06 5.930 58.17 6.029 57.90 5.974 58.02 5.952 57.73 6.023 57.87 6.043 57.57 6.090 57.71 6.052 57.40 5.980 57.56 6.045 57.24 5.930 57.41 6.003 57.08 5.954 57.26 6.057 129 Table A10.5 Data for Figure 3.11 TR TR Concentration Concentration Temperature line Temperature line (°C) 64 7. WM (°C) 64 % wlw 64.94 6.699 65.57 6.965 64.95 6.753 65.54 7.016 64.95 6.801 65.51 6.932 64.95 6.741 65.46 6.925 64.93 6.780 65.40 6.919 64.90 6.683 65.34 7.044 64.87 6.881 65.27 6.957 64.83 6.746 65.19 6.848 64.78 6.636 65.11 6.924 64.72 6.952 65.02 6.859 64.66 6.811 64.92 6.836 64.59 6.796 64.82 6.967 64.51 6.927 64.71 6.878 64.42 7.053 64.60 6.980 64.33 6.696 64.48 6.969 64.24 6.886 64.36 6.960 64.14 6.843 64.24 6.901 64.03 6.680 64.11 6.845 63.92 6.786 63.98 6.945 63.80 6.794 63.85 6.894 63.68 6.811 63.71 6.925 63.55 6.639 63.58 6.968 63.42 6.756 63.44 6.811 63.28 6.719 63.29 6.787 63.15 6.949 63.15 6.916 63.00 6.740 63.01 6.878 62.86 6.806 62.86 6.930 62.71 6.850 62.71 6.929 62.56 6.702 62.56 6.889 62.41 6.756 62.41 6.864 62.25 6.700 62.26 6.903 62.10 6.836 62.11 6.797 61.94 6.983 61.95 6.904 61.78 6.595 61.80 6.838 61 .62 6.964 61.64 6.712 61.46 6.913 61.48 6.791 61 .30 6.568 61.33 6.905 61.14 6.806 61.17 6.865 60.98 6.717 61.01 6.856 60.81 7.370 60.84 6.852 60.65 6.733 60.68 6.848 60.49 6.772 60.51 6.830 60.33 6.526 60.35 6.850 60.17 6.528 60.18 6.722 60.02 6.673 60.00 6.771 59.86 6.711 59.83 6.857 130 Table A10.6 Data for Figure 3.11 TR TR Concentration Concentration Temperature line Temperature line (°C) 65 % wlw (°C) 65 "I. wlw 66.30 7.508 66.34 7.519 66.31 7.425 66.33 7.395 66.31 7.529 66.31 7.526 66.29 7.611 66.28 7.471 66.26 7.501 66.24 7.619 66.22 7.459 66.19 7.524 66.17 7.531 66.14 7.533 66.10 7.495 66.08 7.581 66.03 7.515 66.01 7.440 65.95 7.352 65.93 7.381 65.86 7.533 65.85 7.609 65.76 7.471 65.76 7.535 65.66 7.452 65.67 7.477 65.55 7.537 65.57 7.514 65.44 7.449 65.46 7.401 65.32 7.441 65.35 7.456 65.19 7.414 65.24 7.537 65.07 7.495 65.12 7.501 64.94 7.353 64.99 7.593 64.80 7.349 64.86 7.587 64.67 7.394 64.73 7.483 64.53 7.461 64.59 7.405 64.39 7.475 64.45 7.480 64.24 7.508 64.31 7.465 64.10 7.280 64.17 7.391 63.95 7.504 64.02 7.473 63.81 7.378 63.86 7.507 63.66 7.474 63.71 7.268 63.51 7.275 63.55 7.477 63.36 7.316 63.39 7.323 63.21 7.425 63.23 7.432 63.06 7.273 63.07 7.418 62.90 7.440 62.90 7.339 62.75 7.345 62.74 7.313 62.59 7.296 62.57 7.342 62.43 7.378 62.40 7.381 62.27 7.252 62.23 7.339 62.11 7.345 62.05 7.342 61.95 7.332 61.88 7.344 61.78 7.239 61.71 7.357 61.61 7.354 61.53 7.297 61.44 7.357 61.36 7.395 61.26 7.269 61.18 7.324 61.07 7.295 61.00 7.434 60.89 7.120 60.83 7.348 60.69 7.245 60.65 7.359 131 Table All Data for Figure 4.5 PIR. PIR‘ PIR. Concentration Concentration Concentration Temperature line Temperature line Temperature line °C 35.55 1.810 42.98 2.221 50.33 2.602 35.55 1.821 42.98 2.156 50.33 2.548 35.55 1.808 42.98 2.125 50.33 2.480 35.20 1.797 42.51 2.136 49.84 2.459 34.79 1.827 42.10 2.042 48.88 2.462 34.02 1.748 41.29 2.056 48.42 2.301 33.66 1.751 40.80 2.021 47.91 2.301 32.86 1.746 40.38 2.016 46.96 2.306 32.51 1.723 39.55 1.904 47.91 2.206 38.77 1.871 45.58 2.237 38.38 1.972 44.72 2.194 37.53 1.846 43.82 2.091 43.44 2.123 PlR. PIR. PIR. Concentration Concentration Concentration Temperature line Temperature line Temperature line (°C 59 % wlw (°C 62 96 wlw (°C) 65 it wlw 57.75 3.105 64.65 3.852 67.35 4.200 57.75 3.090 64.65 3.813 67.35 4.099 57.75 3.074 63.91 3.849 66.82 4.043 57.18 3.018 63.31 3.869 66.82 4.015 56.10 2.931 62.63 3.628 65.27 3.833 55.48 2.861 61.95 3.581 64.65 3.628 54.89 2.707 60.75 3.406 63.91 3.674 53.89 2.634 60.13 3.339 62.63 3.420 53.41 2.641 59.00 3.328 61.95 3.388 52.40 2.501 58.32 3.055 60.75 3.362 51.26 2.472 57.18 2.922 50.33 2.454 56.10 2.858 49.84 2.391 55.48 2.806 48.88 2.311 54.46 2.811 53.41 2.618 54.46 2.596 PlR. Temperature PIR1 Solubility (°C) Solubility stdev 31.09 1.696 0.043 36.07 1.835 0.044 40.99 2.027 0.036 45.88 2.231 0.038 50.73 2.461 0.053 55.57 2.805 0.042 60.53 3.253 0.128 65.43 3.778 0.058 132 Table A12 Data for Figure 4.6 PIR. PIR. PIR. Concentration Concentration Concentration Temperature line Temperature line Temperature line 50 %wlw °C 35.55 4.269 42.98 4.990 50.33 6.157 35.55 4.251 42.98 5.031 50.33 6.097 35.55 4.246 42.98 4.927 50.33 5.984 35.20 4.255 42.51 4.930 49.84 6.070 34.79 4.175 42.10 4.962 48.88 6.015 34.02 4.250 41.29 4.886 48.42 5.847 33.66 4.163 40.80 4.938 47.91 5.986 32.86 4.272 40.38 4.998 46.96 5.976 32.51 4.231 39.55 5.047 47.91 5.991 38.77 4.946 45.58 5.917 38.38 4.978 44.72 5.869 37.53 4.892 43.82 5.899 43.44 5.827 PIR. PlR. PIR, Concentration Concentration Concentration Temperature line Temperature line Temperature line °C 59%wlw °C 62%wlw °C 65%wlw 57.75 7.443 64.65 8.669 67.35 10.606 57.75 7.484 64.65 8.863 67.35 10.793 57.75 7.290 63.91 8.526 66.82 10.639 57.18 7.174 63.31 8.435 66.82 10.904 56.10 7.163 62.63 8.808 65.27 10.358 55.48 7.121 61.95 8.709 64.65 10.814 54.89 7.185 60.75 8.509 63.91 10.715 53.89 7.169 60.13 8.831 62.63 10.885 53.41 6.962 59.00 8.450 61.95 10.530 52.40 7.104 58.32 8.661 60.75 10.792 51.26 7.041 57.18 8.735 50.33 7.161 56.10 8.540 49.84 7.160 55.48 8.367 48.88 7.199 54.46 8.588 53.41 8.470 54.46 7.884 54.456 2.596 PIR. Temperature PIR. Solubility °C Solubil etdev 31.09 3.888 0.025 36.07 4.463 0.040 40.99 5.213 0.054 45.88 6.042 0.108 50.73 7.106 0.071 55.57 8.431 0.144 60.53 9.753 0.191 65.43 11.327 0.255 133 Table A13 Data for Figure 4.10 Peak shift Peak shift Peak shift relative to relative to relative to Temperature 5862 nm. Temperature 5862 nm. Temperature 5862 nm. (°C) 50 % wlw (°C) 54 % wlw (°C) 58 % wlw fl 35.55 -2.248 45.58 -4.249 55.48 -5.501 35.55 -2.053 45.58 -3.881 55.48 -5.701 35.55 -2.213 45.17 -3.826 54.89 -5.520 35.20 -2.186 44.72 -3.774 54.46 -5.329 34.79 -2.073 44.32 -3.725 53.89 -5.312 34.02 -1.806 43.82 -3.757 53.41 -5.062 33.66 -1.960 42.98 -3.340 52.40 -4.741 32.86 -1.523 42.51 -2.938 51.86 -4.682 32.51 -2.053 41.65 -2.503 50.82 -4.740 41 .29 -2.981 50.33 -4.575 40.38 -2.961 49.36 -4.474 40.01 -2.370 48.42 -3.824 39.18 -2.695 47.91 -4.114 38.38 -1.953 46.96 -3.862 Peak shift Peak shift Peak shift relative to relative to relative to Temperature 5862 nm. Temperature 5862 nm. Temperature 5862 nm. Solubility (°C) 62 % wlw (°C) 65 % wlw (°C) Solubility stdev 64.65 -8.398 67.35 -8.489 31 .09 -0.977 0.169 64.65 -8.617 67.35 -8.561 36.07 -2.043 0.202 63.91 -8.396 66.82 -8.368 40.99 -2.794 0.170 63.31 -8.330 66.82 -8.415 45.88 -3.838 0.239 62.63 -8.047 65.27 8.477 50.73 -4.749 0.209 61.95 -7.995 64.65 -7.981 55.57 -5.529 0.223 60.75 -7.905 63.91 -7.617 60.53 -6.688 0.228 60.13 -7.179 62.63 -7.243 65.43 -7.756 0.208 59.00 -7.213 61.95 -7.372 58.32 -6.801 60.75 -7.006 57.18 -6.181 56.10 -6.394 55.48 -6.249 54.46 -5.699 53.41 -5.792 54.46 -5.928 134 Table A14 Data for Figure 4.11 Peak shift Peak shift Peak shift relative to relative to relative to Temperature 6996 nm. Temperature 6996 nm. Temperature 6996 nm. (°C) 50 % wlw (°C) 54 % wlw (°C) 58 “I. wlw 35.55 -1.151 45.58 6.328 55.48 9.552 35.55 3.566 45.58 5.331 55.48 7.308 35.55 1.166 45.17 6.405 54.89 8.231 35.20 6.267 44.72 2.868 54.46 7.141 34.79 1.903 44.32 5.754 53.89 8.816 34.02 1.086 43.82 7.244 53.41 9.692 33.66 3.527 42.98 7.188 52.40 7.168 32.86 0.290 42.51 2.859 51.86 6.497 32.51 2.845 41.65 1.680 50.82 5.362 41 .29 6.264 50.33 6.958 40.38 2.811 49.36 6.468 40.01 3.072 48.42 5.502 39.18 1.297 47.91 5.693 38.38 0.036 46.96 4.811 Peak shift Peak shift Peak shift relative to relative to relative to Temperature 6996 nm. Temperature 6996 nm. Temperature 6996 nm. Solubility (°C) 62 % wlw (°C) 65 % wlw (°C) Solubility stdev 64.65 11.394 67.35 12.417 31 .09 1.335 1.9211 64.65 11.345 67.35 12.668 36.07 2.425 1.0928 63.91 12.179 66.82 13.988 40.99 4.618 1.2584 63.31 11.454 66.82 13.071 45.88 5.289 1.2916 62.63 11.045 65.27 12.056 50.73 7.644 1.56 61.95 10.350 64.65 12.310 55.57 8.506 1.2934 60.75 10.703 63.91 9.703 60.53 10.167 1.2655 60.13 10.673 62.63 9.761 65.43 11.315 0.624 59.00 9.926 61.95 12.237 58.32 10.654 60.75 9.477 57.18 9.074 56.10 7.857 55.48 8.644 54.46 7.821 53.41 9.656 54.46 7.131 135 Table A15 Data for Figure 4.12 Peak shift Peak shift Peak shift relative to relative to relative to Temperature 6132 nm. Temperature 6132 nm. Temperature 6132 nm. (°C) 50%wlw 54%wlw 58%wlw 35.55 3.512 45.58 6.967 55.48 8.809 35.55 3.270 45.58 6.081 55.48 9.298 35.55 3.474 45.17 6.719 54.89 8.744 35.20 3.635 44.72 6.585 54.46 9.384 34.79 3.208 44.32 6.328 53.89 8.879 34.02 4.041 43.82 6.019 53.41 9.561 33.66 3.417 42.98 6.549 52.40 8.947 32.86 3.791 42.51 6.761 51.86 8.811 32.51 4.005 41.65 6.757 50.82 9.127 41.29 6.761 50.33 9.322 40.38 6.696 49.36 9.013 40.01 6.841 48.42 8.825 39.18 6.416 47.91 9.564 38.38 7.409 46.96 9.219 Peak shift Peak shift Peak shift relative to relative to relative to Temperature 6132 nm. Temperature 6132 nm. Temperature 6132 nm. Solubility (°C) 62 7. WM (°C) 65 7. WM (°C) Solubility stdev 64.65 10.544 67.35 12.824 31.09 1.393 0.231 64.65 10.705 67.35 12.900 36.07 3.661 0.213 63.91 10.863 66.82 12.952 40.99 5.703 0.095 63.31 10.920 66.82 12.732 45.88 7.477 0.224 62.63 10.782 65.27 13.160 50.73 9.106 0.195 61.95 11.048 64.65 12.741 55.57 10.655 0.233 60.75 11.201 63.91 13.125 60.53 11.975 0.170 60.13 10.920 62.63 13.044 65.43 13.070 0.304 59.00 11.285 61.95 13.320 58.32 11.498 60.75 13.293 57.18 11.163 56.10 11.381 55.48 11.538 54.46 11.974 136 Table A16 Data for Figure 4.13 Peak shift Peak shift Peak shift relative to relative to relative to Temperature 8116 nm. Temperature 8116 nm. Temperature 8116 nm. (°C) 50%wlw 54%wlw 58%wlw 35.55 6.819 45.58 20.491 55.48 38.830 35.55 4.485 45.58 19.741 55.48 38.320 35.55 5.930 45.17 22.158 54.89 37.229 35.20 3.135 44.72 19.635 54.46 39.925 34.79 5.184 44.32 20.274 53.89 39.178 34.02 6.295 43.82 18.960 53.41 36.583 33.66 4.825 42.98 20.863 52.40 38.129 32.86 4.792 42.51 19.793 51.86 35.421 32.51 5.742 41.65 18.537 50.82 34.888 41.29 18.327 50.33 32.548 40.38 15.997 49.36 34.933 40.01 16.935 48.42 34.745 39.18 16.618 47.91 35.662 38.38 17.829 46.96 31.451 Peak shift Peak shift Peak shift relative to relative to relative to Temperature 8116 nm. Temperature 8116 nm. Temperature 8116 nm. Solubility (°C) 62 % wlw (°C) 65 % wlw (°C) Solubility stdev 64.65 53.528 67.35 69.036 31.09 1.131 1.136 64.65 56.172 67.35 68.751 36.07 7.816 1.636 63.91 56.863 66.82 67.874 40.99 17.108 1.437 63.31 56.867 66.82 66.194 45.88 26.476 1.253 62.63 59.180 65.27 65.671 50.73 35.933 1.640 61.95 54.014 64.65 67.824 55.57 46.750 1.564 60.75 54.919 63.91 66.519 60.53 58.742 1.904 60.13 53.250 62.63 67.325 65.43 68.751 2.166 59.00 53.760 61 .95 67.160 58.32 49.954 60.75 64.113 57.18 51.815 56.10 50.560 55.48 49.358 54.46 49.290 53.41 46.604 54.46 46.356 Table A17 Data for Figure 4.17 137 Peak shift Peak shift Peak shift relative to relative to relative to Temperature 3188 nm. Temperature 3188 nm. Temperature 3188 nm. (°C) 50 % wlw (°C) 53 "/o wlw (°C) 56 % wlw 35.55 5.395 42.98 26.090 50.33 39.107 35.55 8.497 42.10 26.010 48.42 42.479 34.02 8.100 40.38 33.027 45.58 42.938 32.51 9.511 37.53 24.272 43.44 40.883 Peak sh? Peak“ Peak shift relative to relative to relative to Temperature 3188 nm. Temperature 3188 nm. Temperature 3188 nm. (°C) 59 "/o wlw (°C) 62 % wlw (°C) 65 % w/w E — — 57.75 60.487 64.65 75.847 67.35 91.172 55.48 59.729 61.95 76.282 65.27 91.234 52.40 62.326 58.32 78.015 62.63 92.522 48.88 60.050 54.46 74.498 60.75 95.901 Peak shifT relative to Temperature 3188 nm. Solubility (°C) Solubility stdev 31.09 1.148 1.361 36.07 17.060 1.901 40.99 32.854 1.604 45.88 47.478 0.577 50.73 60.335 0.954 55.57 70.284 0.990 60.53 82.058 0.212 65.43 92.158 1.026 138 Table A18 Data for Figure 4.18 Temperature (°C) Temperature (°C) Temperature (°C) Temperature (°C) 64.60 0.73 Concentration at solubility at solubility at solubility at solubility (it WM) from Figure 4.12. from Figure 4.13. from Figure 4.17. from Figure 4.6. 50 35.56 34.02 33.10 34.30 54 43.30 42.06 40.60 41.20 58 50.93 50.10 48.90 48.99 62 57.94 57.42 57.93 56.28 65 64.74 64.84 65.25 63.56 Average “meerature (°C) Temperature (°C) stdev 34.25 1.02 41.79 1.17 49.73 0.97 57.39 0.78 Temperature 40.56 40.38 39.39 38.22 37.10 35.78 34.04 32.22 31.47 32.32 32.95 33.07 32.93 32.71 32.20 31.82 31.56 31.25 30.77 30.56 30.46 30.40 30.35 30.33 30.32 30.32 30.31 30.31 30.31 30.00 28.86 27.74 26.79 25.89 24.55 22.88 22.00 21.06 20.43 19.17 18.20 17.17 16.15 15.50 14.57 13.32 12.18 10.47 Table A19 Data for Figure 5.4 139 Temperature 35.30 30.26 25.17 20.05 15.01 9.92 TIR 2.869 2.511 2.229 1.956 1.738 1.468 H ’lv'c" x_"'-.' -xxdli fl MICHIGAN STQTE UNIV. LI illllllilllmiiilillilililiilliMilli! 31293014056‘