t .22 .2 .1135 v31... .9 vs :9. .. 4 5...... ‘11er .35?! v. .. , . 1‘ A In.‘ . I 1:. r: .. . is... :1 C‘J. .. 3. 1:3,; nuns: ‘ .z. ._ a. o... r...“ .. 2 .. >r‘i q!!- . .. ‘ .933 ‘12:...- »llix: \ .‘Ilc , .3 1 1.3.... ‘ I, push} . 31.... TI. 2 . y C .1! . ‘3}. n (V fifitrafi. . . . . . . Qwfiflxum...» 13‘“. ,‘x , I l ‘_l. ' ‘ rt?" CHIGAN STATE UNIVERSITY LIBRARIES Ill“ WWWWillmll‘lll mu 3 1293 01026 9862 l This is to certify that the dissertation entitled NMR INVESTIGATIONS OF CHEMICAL SHIELDING, STRUCTURE, AND MULTIPLE QUANTUM-DYNAMICS OF APATITES presented by Gyunggoo Cho has been accepted towards fulfillment of the requirements for Ph. D degree in Chemistry Major professor / Date &2‘%1923 MS U is an Affirmative Action/Equal Opportunity lnsn'tun'on 0-12771 LIBRARY Michigan State University PLACE IN RETURN BOX to move thII checkout from your record. TO AVOID FINES return on or baton data duo. DATE DUE DATE DUE DATE DUE MSU IoAn Atflnnutlvo ActionlEqunl Opportmlty InstItqun —______Z——-——— NMFI INVESTIGATIONS OF CHEMICAL SHIELDING, STRUCTURE, AND MULTIPLE- QUANTUM DYNAMICS OF APATITES by Gyunggoo Cho A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1 993 ABSTRACT NMR INVESTIGATIONS OF CHEMICAL SHIELDING, STRUCTURE, AND MULTIPLE-QUANTUM DYNAMICS OF APATITES by Gyunggoo Cho The 19F MAS-NMR spectra of a series of fluorapatites, M5F(PO4)3, where M a Ca“, Sr2+ and Ba“, and solid solutions of Ca/Sr fluorapatite have been ob- tained. The crystallographic symmetry about the fluoride ions requires that the chemical shielding tensors be axially symmetric. The principal components of ' the 19F shielding tensor of M5F(PO4)3 are obtained from 19F MAS-NMR spectra using the moments method and Herzfeld and Berger graphical method. The use- fulness of these two methods is demonstrated by using the comparison between experimental spectra and simulated spectra obtained from the chemical shield- ing tensors. The measured chemical shielding tensors enable us to separate the contributions to the Ramsey paramagnetic shielding term from the sigma- and pi- bonding between the alkaline earth metal ions and the fluoride ions. The values of sigma- and pi-bonding contributions to 19F shielding for M5F(PO4)3 (M - Ca2+, Sr’+ and 8a“) are 81.7 ppm and 24.6 ppm, 97.6 ppm and 26.0 ppm, and 138.4 ppm and 32.1 ppm respectively with respect to free fluoride ion. The preference of Sr“ ions for the Ca(2) site for fluorapatite have been studied using 19F MAS-NMR spectra of a solid solution of composition Cag,97$r1,03F2(PO4)6. The peak intensities obtained from the centerband and sidebands as well as the deconvolution peak indicate that Sr2+ ions have a 23% preference for the Ca(2) site, which is adjacent to the fluoride ion. The assignment of the spectra of Ca3,978r1.03F2(PO4)5 is aided by the existence of spin diffusion performed by the SPARTAN pulse sequence. The dimensionality of the distribution of spins in solids influences their multiple- quantum NMR dynamics. We have studied these dynamics for the quasi- one-dimensional distribution of uniformly spaced proton spins in hydroxyapatite, Ca5(OH)(PO4)3, and related compounds, using a phase-incremented even order selective MQ pulse sequence. The increase in effective size N with prepara- tion times for stoichiometric monoclinic hydroxyapatite is linear at early times, in agreement with calculations based on the incremental shell model; however, the experimental slope is three times greater than the predicted slope. An upward curvature observed at longer preparation times is qualitatively ascribed to the in- complete isolation of linear chains. A slight deficiency of hydroxyl groups in a sample of hydroxyapatite in the commonly-occurring hexagonal crystal form leads to a measurable decrease in the slope of the linear portion of the curve. The 1H multiple-quantum dynamics of a series of fluorohydroxyapatite solid solutions, Ca5(OH)xF1-x(PO4)3, exhibit decreased slopes for lower hydroxyl levels (smaller x), and requires consideration of the different lengths of spin “clusters” in order to model the behavior. The defect densities of apatites (clusters) are estimated by using 1-D cluster model. We have also studied the 19F multiple-quantum NMR of a single crystal of mineral fluorapatite at different orientations with respect to the external magnetic field. The observed oscillatory behavior of the multiple-quantum dynamics is interpreted in terms of 1-D clusters of fluoride ions in the defect-containing sample. To my mother, wife, and son ACKNOWLEDGMENTS I would like to sincerely thank Dr. James P. Yesinowski for his continuous guidance and enthusiastic encouragement throughout my research. I would also like to thank Dr. Dye, Dr. Cukier, and Dr. Schwendeman for helpful discussions. My special thanks go to Liam B. Moran for helpful discussions of my research. I would like to thank Mr. Kermit Johnson for helpful discussions of new NMR techniques and aid with computers. Thanks are also extended to the members of the NMR facility and the members of machine shop, in particular to Dr. L. Le, Mr. D. Jablonski, and R. Geyer. I would also like to acknowledge GTE Electrical Products Corporation for partial financial support of this work. i would also liked to thanks Dr. Chung-Nin Chau of GTE Electrical Products Corp. for preparing the fluorapatite samples and Dr. B. Fowler of National Institute of Standards and National Institute of Dental Research for lending us the samples of monoclinic hydroxyapatite and single-crystal fluorapatite. I would like to thank my Korean friends at Michigan State University who encouraged me when l was discouraged. Finally, my special thanks are reserved for my family: my wife, Sukhee, my son, Nammyoung, and my mother, who have prayed and waited patiently for my success. vi TABLE OF CONTENTS LIST OF TABLES ..................................... xi LIST OF FIGURES ................................... xii I. INTRODUCTION TO APATITE STRUCTURE AND CHEMISTRY. 1 References ..................................... 6 ll. PART 1 19F MAS-NMR Investigation of Alkaline Earth Fluorapatltes: Measurement of Chemical Shielding Tensors and Characteriz- ation of Sites In Ca/Sr Fluorapatlte Solid Solutions 1. Introduction ................................... 9 2. Nuclear Spin Interactions in Solids .................. 11 A. Zeeman Interaction ............................. 11 B. Radiofrequency Interaction ........................ 12 C. Chemical Shielding Interaction ...................... 12 D. Dipolar Interaction .............................. 14 3. Magic Angle Spinning (MAS) ....................... 17 4. Methods for Measuring the Chemical Shift Anisotropy from MAS-NMR Spectra .......................... 20 A. Moments Method .............................. 20 B. Herzfeld and Berger Graphical Method ................. 21 5. Experimental .................................. 22 A. MAS-NMR Studies ............................. 22 B. Simulation of 19F MAS-NMR Spectra of Fluorapatlte Samples . . . 24 6. Results ...................................... 25 vii A. Alkaline Earth Fluorapatites ........................ 25 B. Solid Solutions of Ca/Sr Fluorapatite .................. 34 7. Discussion ................................... 42 A. Measurement of the 19F Chemical Shift Anisotropies of M5F(PO4)3 Using Different Methods .......................... 42 B. Separation of Metal-Fluoride Sigma- and Pi-Bonding Contributions to the 19F Shielding Tensor ........................ 44 C. Study of Site-Preference of Sr2+ in Ca/Sr Fluorapatite Solid Solutions Using 19F MAS-NMR ....................... 53 8. Conclusions ................................... 63 9. References ................................... 65 III. PART 2 1H and ”F Multiple-Quantum NMR Dynamics of Quasi-One- Dlmenslonal Spin Distributions In Apatltes 1. Introduction .................................. 69 2. Multiple-Quantum NMR Dynamics .................. 75 A. Density Operator Description of Multiple-Quantum NMR Dynamics ................................... 75 B. Time Development of the Density Operator in the Rotating Frame ..................................... 79 C. Pulse Sequence for Multiple-Quantum NMR .............. 81 D. Statistical Model of MO Coherence Intensities ............ 91 E. Simplified Models of Multiple-Quantum Dynamics .......... 92 a. Hopping Model .............................. 95 b. Incremental Shell Model ....................... 104 viii 3. Experimental ................................. 109 A. Multiple-Quantum NMR Studies .................... 109 B. Sample Preparation and Characterization ............... 113 4. Results ..................................... 114 A. 1H Multiple-Quantum NMR Study of Resonance-Offset Effects. . 114 B. 1H Multiple-Quantum NMR of Hydroxyapatite and Fluorapatite Samples .................................. 121 C. 19F Multiple-Quantum NMR of a Single Crystal of Fluorapatite . 131 5. Discussion ................................... 131 A. Effect of Resonance Offsets in PI-MO NMR Pulse Sequence on the Formation of Multiple-Quantum Coherence ......... 131 B. Dimensionality Effects in the Multiple-Quantum NMR of hydroxyapatite ............................... 1 34 C. One-Dimensional Cluster Model .................... 136 D. Estimation of Defect Densities Using the 1-D Cluster Model . . . 139 E. 19F Multiple-Quantum NMR Dynamics of Single Crystal Fluorapatite ................................. 1 44 6. Conclusions .................................. 149 7. References ................................... 154 Appendix A. Phase-incremented multiple-quantum pulse sequence program for solids (Varian VXR 400 spectrometer) ..................... 163 Appendix 8. Selection rule of the 1-spin/2-quantum average Hamiltonian ........................... 167 Appendix C. C program for calculation of the ratio of 40/20 intensities for 1-D chain with different defect densities ............................ 1 70 Table 1. Table 2. Table 3. Table 4. Table 5. Table 6. LIST OF TABLES 19F NMR chemical shift anisotropies and asymmetry parameters of M5F(PO4)3 calculated using the moments method .......... 26 The 19F chemical shift anisotropy and asymmetry parameter of M5F(PO4)3 obtained using the Herzfeld and Berger graphical method ............................... 32 19F chemical shielding tensors of M5F(PO4)3, calculated from the moments method using peak intensities obtained from deconvolution ............................. 32 19F NMR paramagnetic shielding parameters a; and an for M5F(PO4)3 .................................. 51 Calculated probabilities of configurations A-D in Ca/Sr solid solution . 58 Calculated probabilities of configurations H" in Ca/Sr solid solution . 58 LISTOF FIGURES Figure 1. Structure of hydroxyapatite .......................... 5 Figure l-1. Chemical shielding powder pattern .................... 13 Figure l-2. Cartesian axis system and polar coordinates for the dipolar coupling of two nuclei ........................... 16 Figure l-3. Orientation of magnetic field and spinning axis vectors for magic angle spinning experiments ........................ 18 Figure l-4. SPARTAN pulse sequence for monitoring spectral spin diffusion . . 23 Figure l-5. Contour plots of Ca5F(PO4)3 spinning at 6.12 kHz and 10.18 kHz using the Herzfeld and Berger graphic method ............ 29 Figure l-6. Contour plots of Sr5F(PO4)3 spinning at 6.26 kHz and 10.20 kHz using the Herzfeld and Berger graphic method ............ 30 Figure l-7. Contour plots of Ba5F(PO4)3 spinning at 6.23 kHz and 10.18 kHz 1 using the Herzfeld and Berger graphic method ............ 31 Figure l-8. The experimental and simulated spectra of M5F(PO4)3 ....... 35 Figure I-9. 19F MAS-NMR center band HHLW vs spinning speed for M5F(PO4)3 .................................. 36 Figure l-10 ‘9F MAS-NMR of Ca3,978r1,03F2(PO4)5 spinning at 8.23 kHz and at 10.80 kHz ................................. 38 Figure l-11. Deconvolution spectrum of the 64 ppm peak in Fig. l-10b ..... 39 Figure l-12. 19F MAS-NMR spin diffusion experiment of Ca3,978r1,03F2(PO4)6 using the SPARTAN pulse sequence and spinning at 10.87 kHz . 40 Figure I-13. 19F MAS-NMR of CaSSr5F2(PO4)5 spinning at 8.25 kHz and at 10.3 kHz ................................. 41 xii Figure l-14. Figure l-15. Figure l-16. Figure l-17. Figure l-18. Figure ”-1. Figure "-2. Figure "-3. Figure ll-4. Figure "-5. Figure "-6. Comparison of the simulated and experimental 19F MAS-NMR stick spectra of M5F(PO4)3 at the same spinning speed ..... 45 a0 and an for M5F(PO4)3 vs. 19F isotropic chemical shift with respect to free F' ion ........................ 52 Substitution of Sr2+ in the three nearest neighbor Ca(2) sites of Ca/Sr fluorapatite solid solutions .................... 54 Possible configurations for single and double Sr2+ substitution in the six next-nearest neighbor Ca(2) sites of Ca/Sr solid solutions of fluorapatite .......................... 57 Variation in the instantaneous chemical shifts during a rotor cycle of 64 ppm and 79 ppm peaks of the solid solution under the MAS condition at the different crystallite orientations . . 62 Schematic idealized arrangement of linear columns of protons in calcium hydroxyapatite (Ca50H(PO4)3) ............... 73 Random phases and correlated phases in an ensemble of two level systems .............................. 77 General form of the pulse sequence of MO NMR experiments . . . 82 Pulse sequence for multiple-quantum NMR in solids ......... 85 Time-domain MO interferogram of hexamethylbenzene using the TPPl ................................ 86 Frequency-domain 1H MO spectrum of hexamethylbenzene using the TPPI ................................ e7 xiii Figure “-7. Time-domain MO interferogram of hexamethylbenzene using the phase-incremented method .................. 89 Figure "-8. Frequency-domain 1H MO spectrum of hexamethylbenzene using the phase-incremented method .................. 90 Figure ”-9. Schematic energy level diagram for an N (odd) spin-1/2 system in a Zeeman field, and the degeneracy number of each state . . . 93 Figure ”-10. Symbolic representation of the spreading of multiple spin corre- lations in a coupling network with increasing preparation time . . . 94 Figure "-11. Projection of Liouville space onto a two-dimensional plane . . . . 97 Figure "-12. Pathways of the growth of MO coherences in Liouville space for a 6-spin system under the 1-spin/2-quantum Hamiltonian . . . 98 Figure "-13. Evolution of n-quantum coherences predicted by the hopping model in 6 and 20 spin systems ..................... 102 Figure "-14. Development of effective size with increasing preparation times for a 20 spin system predicted by the hopping model ....... 103 Figure "-15. Diagram representing the average M spin operator and its constituent operators of M a 4 in a 9 spin system ......... 106 Figure "-16. Schematic diagram of the growth of MO coherence of different dimensionalities .............................. 1 10 Figure "-17. Values of effective size N predicted by the incremental shell model for spin-1/2 nuclei evenly spaced on a line (1-D), square (2-0). and cubic (3-0) lattice ...................... 111 xiv Figure "-18. Figure "-19. Figure "-20. Figure "-21. Figure “-22. Figure "-23. Figure "-24. Figure "-25. Figure "-26. Figure "-27. Figure "-28. Figure "-29. 1H Pl-MO-NMR spectra of hexamethylbenzene with different transmitter carrier frequencies ..................... 115 1H 1-dimensional spectra of apatite samples ............ 116 1H MO NMR spectrum of HAP-M using a preparation time of 864 [LS ................................. 118 Intensities of the 1H Pl NMR MO order of HAP-M for various resonance offsets using a preparation time 01864 [IS ....... 119 Intensities of the 1H Pl NMR MO order of HAP-M for various resonance offsets using a preparation time of 1872 p8 ...... 120 The normalized intensities of even-order MO coherences of apatite samples as a function of preparation time .......... 122 Effective size N vs. MO preparation time for HAP-M ........ 125 In(N) vs. In(r/rc) for all of the samples of hydroxyapatite and fluorohydroxyapatite ........................... 1 26 Decay of absolute MO coherences of apatites ........... 128 The normalized intensities of OO and ZO coherences of a single crystal mineral sample of FAP with preparation times at two orientations ............................. 129 In(absolute intensity of MO orders) vs. preparation time of single crystal fluorapatite at two different orientation ....... 130 The effective magnetic field in the rotating frame in the presence of an applied rf field .......................... 132 XV Figure ”-30. Figure "-31. Figure "-32. Figure "-33. Figure "-34. Figure "-35. Figure "-36. Arrangement of hydroxyl groups separated by vacancies and/or fluoride ion substitutions in apatite samples ............. 138 Run number distributions for various defect densities in a one- dimensional chain . . . .. ......................... 141 1-D cluster model of MO growth for various defect densities in apatite, compared to experimental data ............... 142 Comparison between experimental 40/ZO intensities and calculated 40/ZO intensities for FOHAP samples after excluding two hydroxyl groups adjacent to fluorine ions ...... 143 Simulated MO dynamics of 2, 3, 4, and 5 spin systems under ”Hg, for an oriented linear chain using ANTIOPE .......... 145 Mole fraction of spins in various run numbers and sum of the mole fraction with different defect densities using percolation theory .................................... 150 Comparison between experimental 20 intensities of single crystal FAP and calculated 20 intensities of two assumed defect densities at two different orientations ........................ 151 xvi I. INTRODUCTION TO APATITE STRUCTURE AND CHEMISTRY Although the chemical reactions leading to mineral formation in biological systems are not fully understood, it is believed that a nonstoichiometric ”defect" form of calcium hydroxyapatite, Ca5(OH)(PO4)3, is the primary mineral phase of bone, dentin and dental enamel.” Apatites' are also important in the production of fertilizer, in the lighting industry as a phosphor in fluorescent lamps, and in chromatography.‘ At the pH values typically found in biological systems, the “stability of calcium phosphates increases and the solubility decreases as the molar Ca/P ratio increases. Thus, in vivo dicalcium phosphate dihydrate CaH PO4-2H20 (DCPD, CalP=1.00) hydrolyzes into octacalcium phosphate Ca3H2(PO4)6-5H20 (OCP, Ca/P=1.33), which hydrolyzes into hydroxyapatite (HAP) (Ca/P=1.67).5 Although amorphous calcium phosphate (ACP) is not found in bone and teeth, it can occur in vivo and is transformed into crystalline apatite via an octacalcium phosphate-like phase.6 The substitution of numerous impurities in the apatite lattice changes the properties of apatites. For example, the fluoride ion is effective in prevents the dental caries, since it increases the rate of remineralization and lessens the acid demineralization.“ The surface of HAP reacts with fluoride ions to yield calcium difluoride (Can), fluorohydroxyapatite (Ca5(OH)1-xe(PO4)3), and fluorapatite (Ca5F(PO4)3), depending upon conditions. Although the fluoridation of hydroxyapatite is not completely understood, several mechanisms have been suggested, including the ionic exchange of F' for OH‘ in the apatite structure,9 direct precipitation of fluorapatite mineral, and dissolution of hydroxyapatite and recrystallization of fluorapatite in the presence of F'.‘° Carbonate ions can also replace hydroxyl groups or phosphate groups in the apatite lattice. A carbonate ion 1 can substitute for two hydroxyl groups in hydroxyapatite, forming type A carbonate apatite. This reaction takes place at high temperature,11 and results in an increase of the a axis length.12 The substitution of phosphate groups by carbonate ions decreases the a axis and increases the c axis.“3114 The product is referred to as type B carbonate apatite.11 Calcium ions in apatites can be replaced by strontium ions. The presence of strontium ions in bone and teeth increases susceptibility to caries. The incorporation of strontium ions increases the a axis and c axis length.15 From X-ray crystallographic analysis, hydroxyapatites have either a hexagonal crystal system with the space group P63/m or, for very stoichiometric samples, a monoclinic system with the space group P21/b. In the hexagonal crystal system, there are two Ca5(OH)(PO4)3 groups in a unit cell of dimensions a (=b) = 942 pm and c = 688 pm.5 The structure of hexagonal hydroxyapatite is shown in Figure 1 projected down the c axis. There are two types of calcium ions in the structure, Ca(1) and Ca(2). Hydroxyl groups are surrounded by three Ca(2) ions in an equilateral triangle. Infinite linear chains of protons have a uniform spacing of 344 pm. The position of the hydroxyl protons is located about 130 pm from the plane of the triangle of Ca(2) ions. The central column of hydroxyl groups has six hexagonalIy-situated neighboring columns at distances of 942 pm. Statistical disorder of the hydroxyl groups result, on average, in protons in three of these six columns being located about 130 pm below the Ca(2) triangle, and about 130 pm above the Ca(2) triangle in the omer three columns. When viewed down the (hexagonal) c axis, the Ca(2) ions form two displaced equilateral triangles of Ca(2) ions which are rotated by 60°. The distance between the two triangles is 344 pm. Columns of Ca(1) ions are parallel to the c axis. Each Ca(1) column is located at the middle of a large equilateral triangle of three hydroxyl groups. The unit cell dimensions of monoclinic apatite are a = 942 pm, b = 2a, c = 688 pm, and 7 = 120°.‘6'17 The space group of monoclinic hydroxyapatite, P21/b, indicates that the monoclinic form is regarded as twins occurring at 120° rotations about the c axis.“5v17 In the monoclinic crystal form, the protons in four of the six neighboring columns are located about 130 pm below the plane of the three Ca(2) ions, and those in the other two columns are located about 130 pm above the plane of the three Ca(2) ions. Fluorapatite, which has a structure similar to that of hydroxyapatite, has the same unit cell dimensions except that a = 937 pm. Fluoride ions are exactly in the middle of the plane of the three Ca(2) ions, which forms a mirror plane.18 Apatites have been investigated using various methods such as X-ray crys- tallography, lR, Raman, NMR, etc. The chemical environment of fluorine ions in alkaline earth fluorapatites and the preference of 812+ ions for Ca(2) sites of Sr/Ca fluorapatite solids solutions are studied using 19F MAS-NMR. The high level of defects found in naturally-occurring apatites is of interest in its own right, since they presumably reflect the conditions of formation of the mineral. Fluoride ions of fluorapatite are interrupted by charge-coupled vacancies, or substitution of fluoride ions by carbonate groups, hydroxyl groups, or chlorine ions. Hydroxyl chains in hydroxyapatite also have vacancies, or hydroxyl group substitutions by carbonate groups, fluorine ions, or chloride ions. Vacancies of hydroxyl groups in hydrox- yapatite have been quantitatively determined by using IR and 1H MAS-NMR.20 However, these methods cannot reveal the presence of microscopic heterogene- ity of various types of anions. The 1H and 19F multiple-quantum NMR experiments in this thesis provide for the defect in the hydroxyl or fluorine chains and make it possible to calculate the defect densities of apatites. Furthermore, the quasi- one-dimensional spin distributions in apatites represent a valuable model system in which to study the effects of dimensionality upon multiple-quantum NMR dy- namics. 11a. o 3:, 3,) a Figure 1. Structure of hydroxyapatite. Calcium ions (large open circle), hy- droxyl groups (dotted circle), and phosphorus group (small open circles connected by a line). See text. Taken from Ref. 5. References 1. W. E. Brown, Clin. Orthop. 44, 205 (1966). 2. D. G. A. Nelson, J. D. B. Featherstone, Calcif Tissue Int. 34, (Suppl.2) 869 (1982). 3. R. Z. LeGeros and J. P. LeGeros, in ”Phosphate Minerals” (J. O. Niragu and P. B. Moore, Eds), pp. 351-385, Springer-Verlag, Berlin, 1984. 4. S. J. Joris and C. H. Amberg, J. Phys. Chem. 75, 3167 (1971) 5. W. E. Brown, M. Mathew, and M. S. Tung, Prog. Cryst. Growth Charact. 4, 59 (1991). 6. J. L. Meyer and C. C. Weatherall, J. Colloid Sci, 89, 257 (1982). 7. L M. Silverstone, Caries Res. 11, 59 (1977). 8. D. J. White, Caries Res. 22, 27 (1988). 9. S. A. Leach, Brit. Dent. J. 106, 133 (1959) 10. M. A. Spinelli, G. Brudevold, and E. C. Moreno, Arch. Oral Biol. 16, 187 (1971). 11. G. Bonel and G. Montel, CR Acad. Sci. [0] (Paris) 258, 923 (1964). 12. J. C. Elliott, " The crystallographic structure of dental enamel and related apatites", Ph.D. Thesis, University of London, 1964. 13. R. Z. LeGeros, Nature, 206, 403 (1965). 14. R. Z. LeGeros, J.P. LeGeros, O.R. Trautz, and E. Klein, J. Dent. Res. 43, 751 (1964). 15. H. J. M. Heijigers, F. C. M. Driessen, and R. M. H. Verbeeck, Calcif Tissue Int. 29, 127 (1979). 16. J. C. Elliott, Nature Phys. Sci, 230, 72 (1971). 17. J. C. Elliot, P. E. Mackie, and R. A. Young, Science, 180, 1055 (1973). 18. K. Sudarsanan, P. E. Mackie and R. A. Young, Mater. Res. Bull, 7, 1331 (1972). 19. E. C. Moreno, M. Kresak, and R. T. Zaradnik, Caries Res. Supp/.1 142 (1977) 20. J. Arends, J. Christofferson, M. R. Christofferson, H. Eckert, B. O. Fowler, J. C. Heughebaert, G. H. Nancollas, J. P. Yesinowski and S. J. Zawacki, J. Cryst. Growth, 84, 515 (1997). PART 1 19F MAS-NMR investigation or Alkaline Earth Fluorapatltes: Measurement of Chemical Shielding Tensors and Characterization of Sites in CalSr Fluorapatlte Solid Solutions 1. Introduction The chemical shielding tensors of 19F in metal fluoride salts have been of widespread experimental and theoretical interest."5 Line narrowing techniques such as multipulse “dipolar decoupling,"2 magic-angle spinning (MAS),6'7 or the combination of these two techniques (CRAMPS)8 have been employed to obtain more accurate values for the isotropic chemical shift. However, knowledge of the three components of an anisotropic chemical shielding tensor can provide a more detailed understanding of the factors governing the chemical shielding, if the structural environment of the fluorine atom is known. Although single-crystal NMR studies are generally required to orient an arbitrary chemical shielding tensor in the crystallographic axis system, if the shielding tensor is axially-symmetric, one of its principal axes must necessarily be along the corresponding symmetry axis of the crystal. The fluoride ion in the alkaline earth fluorapatites (M5F(PO4)3, M = Ca, Sr, Ba) resides on a crystallographic hexagonal screw axis,"'11 and therefore possesses such an axially symmetric 19F chemical shielding tensor. The 19F chemical shift anisotropy (CSA) for calcium fluorapatite has been determined from single- crystal NMR measurements”:13 and an accurate isotropic chemical shift has been obtained from 19F MAS-NMR studies (because of the special arrangement of the fluoride ions in a linear chain, sharp spectra are obtained at modest spinning speeds without using multiple-pulse techniques, as discussed in references 6 and 7). However, corresponding data for the strontium and barium analogues have not been reported. We have therefore measured the 19F chemical shielding tensors of these compounds, using three different methods that yield the CSA from the spinning sideband patterns (spectral moments,” Herzfeld-Berger graphical 10 analysis,15 and spectral simulation), and will discuss briefly the relative merits of these methods. Knowing both the chemical shielding tensor components and the absolute chemical shielding scale relative to the free fluoride ion enables one to separate contributions to the shielding from the sigma— and the pi-bonding between metal and fluoride ion for the three different metals investigated. This approach enables us to predict chemical shielding tensors in other systems, such as solid solutions of calcium/strontium fluorapatite. The broad linewidth typical of powder samples arising from dipolar interactions and CSA obscures identification of individual peaks having different chemical environments in many solids of interest. Since MAS-NMR averages out these anisotropies, and results in sharp peaks at each isotropic chemical shift, it is - extremely useful for the structural study of solids having many components. For example, the site preference of solid solutions of a semiconductor alloy was studied by using 31P MAS-NMR and peak deconvolution,16 and ”F MAS-NMR has been also used to quantitatively studying fluoride ions perturbed by antimony ions in antimony-doped fluorapatites.20 Since the chemical shift of fluorine is very sensitive to its chemical bonding environment, we have used high speed 19F MAS-NMR for the study of solid solutions of calcium/strontium fluorapatite. The 19F NMR parameters measured for the pure alkaline earth fluorapatites (M5F(PO4)3, M = Ca, Sr, Ba) are used in the interpretation of the data from the solid solutions. Since the 19F chemical shift difference between strontium and calcium fluorapatites is 33.2 ppm, the isotropic chemical shifts of samples having different ratios of calcium to strontium can be distinguished by using high field/high speed 19F MAS-NMR. The integrated intensity of the non-overlapped peaks and 11 the deconvolution of asymmetric broad peaks arising from the perturbation by strontium substitution yield the ratios of each component. The results were used to study the site preference of Sr2+ ion substitution in a solid solution having the composition Cag,gysr1,o3F2(PO4)5. 2. Nuclear Spin Interactions in Solids A typical nuclear spin Hamiltonian of a diamagnetic solid is given by ”H: H2+Hrf(t)+HCS+HD‘I-HQ+HJ (M) where the various Hamiltonians on the right represent respectively the Zeeman, radiofrequency, chemical shift, dipolar, quadrupolar, and scalar interactions of the nuclei. The first two terms are determined by external static magnetic and applied rf fields. Thus, they describe “external” interactions of the spins. The other terms depend on the fundamental characteristics of the nucleus and its environment. In the solids that we have studied, the quadrupolar and scalar interaction are not relevant, and will not be discussed. A. Zeeman Interaction The interaction between a nuclear magnetic moment, it, and an applied static field, 130 is represented by the Zeeman Hamiltonian H, = -§L;Zflt'fio= -Z'rih1;;'fio 0'2) where h is Planck’s constant divided by’21r, 7,- is the magnetogyric ratio of nucleus i, and 1‘; is the angular momentum spin operator for nucleus i. The eigenvalues of this term alone are Ez/h = 2:.7IhmiHo 0‘3) where m,- is the magnetic quantum number. 12 B. Radiofrequency Interaction In NMR spectroscopy, transitions between energy levels are generally induced by an applied rf field, which is applied perpendicular to the static field direction. The Hamiltonian term for the radiofrequency (rf) field along the x direction is given by H,f(t) = 27rjl'3H1COS(2m/I) (l-4) where ii, is the nuclear magnetic moment in the x direction, H1 is the magnitude of the if field applied in the x direction, and u is the Larmor frequency. C. Chemical Shielding Interaction The screening of the nuclei from the external magnetic field by the surrounding electrons slightly modifies the Zeeman interaction. The shielding generated at the nuclei from the external magnetic field results in the Hamiltonian ”HOS .-. vilified, (15) where 6 is the chemical shielding tensor, a dimensionless second rank tensor. In solutions, the chemical shielding interactions are averaged out by rapid isotropic tumbling. Thus, a single line in solution is observed at the isotropic average of the shielding tensor (Tr{&}). In solids, since molecular motions are typically slow or absent, a broad powder pattern is observed. The chemical shielding tensor 6 is symmetric in larger static magnetic field. In the principal axis system (PAS), all off-diagonal elements of shielding tensor zero. The chemical shielding tensor can ne described by the three principal values 011, 022, and 033, and three angles which specify the orientation of the principal axis system. If ”as are the dominant interactions, the three components of the chemical shielding tensor can be “read- off' directly from the spectrum since a powder pattern is related the chemical shielding. The theoretical powder patterns for a shielding tensor with different I3 bl ' I 0'39 ., - / 03 ———J .\_h ‘ 'r'*~'i "r" " i I 140130120110 :00 9C 30 70 60 50 4C BC 20 ppm Figure l-1. Calculated chemical shielding powder patterns (a,- =- 64 ppm): (a) axially symmetric shielding tensor (CSA = 84 ppm and i) a 0) ; (b) non-axially symmetric shielding tensor (CSA :1: 84 ppm and n = 0.3). The dashed line denotes the isotropic chemical shift An exponential apodization function corresponding to a 500 Hz line broadening was applied to the calculated FID before Fourier transformation. 14 asymmetry parameters 77 are shown in Fig. H. The asymmetry parameter 7) is defined as n = FL“ (l-6) Conventionally, the order of the principle values of the chemical shielding tensors is 011 s 022 s 033. The chemical shift anisotropy is defined as CSA = 033 - 1/2(011 + 022). (1-7) The chemical shielding tensor is generally obtained from single crystal NMR studies or as in the present work, from MAS-NMR spectra on polycrystalline samples. D. Dipolar Interaction The dipolar interaction is the consequence of direct magnetic coupling of nuclei through space. The dipolar Hamiltonian of two nuclear spins (spin 1 and spin 2) is represented by ”D: (Vi/Thin? it 'iji M) where r12 is distance between the nuclei and D12 is the dipolar coupling tensor. In a Cartesian coordinate system x, y, 2 (see Fig. I-2), (rfi - 3x2) —3xy -3xz . D12 = l/rfi —3xy (1%? — 3y?) —3yz (I-9) —3xz —3yz (rfi — 3Z2) The trace of D12 is zero, and it is axially symmetric: D13 = 021, 023 = D32, and 0,3 = 03,. In the principal axis system, with r12 along the z axis, all off-diagonal elements are zero. The dipolar tensor can then be rewritten as . 1 0 0 D=1/ri’2 0 1 0 (HO) 0 0 —-2 It is useful to transform from Cartesian coordinates to polar coordinates (Fig. l-2). 15 The Cartesian coordinates x, y, and z are represented by the following x = rsin 6 cos d) y=rsin09in¢ (H1) 2 = r cos 6. Eq. (l-8) can be rewritten in the polar coordinate system as HD==7/rf,h(A-i-B+C+D+E+F) 0'12) A = (1 — 3CO529)IlzI2z 9:...)(1 _ 3cos 2a) [ITIg + 171;] = 3(1 — aces 29) (11.12. — f1 - f2) = _gsin9cos9e-‘¢ [1,ng +IT122] (I-13) D = —-:2isin9c090t=3'm5 [11.212 +ITI22] E =—%sin266'2i¢ITIl§ = —%sin29e+2‘¢I;I; where I + and I ‘ are the ladder operators, I + = f; + if; and I “ = I; — if, I+|a) = 0, I+|fl) = Id), I'|a) = Id), and I‘lfl) = 0. The various ladder operators can change the nuclear spin quantum numbers if), and mg in a characteristic way. Term ’A' does not shift the nuclear spin quantum number; term '3’ alters both spins by t 1, but I TI 3 and I II *2" do not change the sum (11), + mg); the others change (m + mg) by :t 1 or :I: 2. Since, at high magnetic fields, the perturbations due to the off-diagonal elements C, D, E, and F are negligible compared to the terms ’A' and '8’, Eq. (l-13) can be rewritten as the truncated homonuclear dipolar Hamiltonian in terms of the A and B terms 16 Figure l-2. Cartesian axis system and polar coordinates for the dipolar coupling of the two nuclei. 17 H0 = (ii/2ri.)n(1— 3cos 26001-11; — 311.12.). (l-14) The truncated heteronuclear dipolar Hamiltonian remains only the term ’A' that is replaced It and [2 by I and S. 3. Magic Angle Spinning (MAS) In solids, anisotropic interactions such as the chemical shielding interaction, dipolar couplings and electric quadrupole couplings can result in very broad NMR spectra. These anisotropies can be efficiently removed by spinning the sample about an axis making an angle of approximately 54.7° with respect to the external magnetic field”19 (Fig. l-3) (see below). The rotation of the sample about an angle 9 with respect to the external magnetic field makes the anisotropic terms of the Hamiltonian time-dependent with the periodicity of the sample rotation frequency 1.2,. The three anisotropic terms mentioned above are generally small compared to a Zeeman term and are treated as perturbations. The Hamiltonian is divided into two parts H=H+Hh) am) where 72 is the time-average of the Hamiltonian and ’H'(t) is time-dependent. In this section, the effect of sample rotation on each of two anisotropic terms will be discussed in turn. Since the chemical shielding tensor is of small size relative to the Zeeman term, it can be truncate it. We can rewrite Eq. (l-5) as He, = 7hazzH0 (I-16) where 0n = M21011+ A222022 + M23233- ("17) 18 Figure l-3. Orientation of magnetic field Ho and spinning axis vectors for magic angle spinning experiments. 19 In these equation, the terms Uaa (a = 1, 2, 3) are the principal values of the chemical shielding tensor and AW are the direction cosines of the principal axes with respect to the external magnetic field. The rotation of the sample makes the direction cosines time-dependent A, = cos 8 cos X, + sin 8 sin X, cos (wrt + 8,) (I-18) where ,8 is the angle between the rotation axis and the external magnetic field, X, is the angle between the rotation axis and the p-th principal axis of the chemical shielding tensor, and 18,, is the azimuthal angle of the p-th principal axis at t=0. By substituting Eq. (I-18) into Eq. (H7) and taking the time-average, we obtain the following equation as: = %Sin2fiTr{&} + %(3cos2i’3 — 1) 20,, cosxp. (I-19) Only the isotropic chemical shift 0,- (=1 /3Tr{ 6}) in Eq. (l-19) remains if [3 is 54.7° [(3 cos28 — 1) = 0], the so-called “magic angle”. Since the rotation of the solid sample makes cos 6 in the dipolar Hamiltonian time-dependent, we can rewrite cos 6 as cos 6(t) = cos 8 cos 8' + sin 8 sin 8’ cos (8.1 + 8’) (l-20) where 8' is the angle between the axis of rotation and F12, and this the azimuthal angle of £12 at 1:0. The time average of cos2 0(t) is 2637? = cos28cos28’ + %sin2fisin2 8' = F§.(3cos2 8 — 1) (3cos2 8’ -— 1) + I, (I-21) Substituting Eq. (I-21) into Eq. (l-14), we can write the average (time-independent) dipolar Hamiltonian as Hp = guzcos2 8 — 1) (3 cos2 8’ — 1) (i, . i; — 311.12,). (1-22) The time-independent dipolar interaction is averaged out at the magic angle. Under the MAS condition, the averaged time-independent Hamiltonian yields only isotropic chemical shifts, but time-dependent Hamiltonians are modulated by 20 the periodicity of the Spinning speed of a sample. From Eq. (H4) and (l-22), the time-dependent part of the dipolar interaction is given by Hm) = (37f/2r3)h(f1 - is -— 31,12) [sin 28 sin 28’ cos (wrt + 8) + sin2 8 sin2 8’ cos2(.c,.t + 18) (l-23) The time-dependent Hamiltonians of the other two interactions (chemical shift and quadrupolar interactions) also show the same periodicity in w. and 28:, as the time—dependent dipolar Hamiltonian. The modulations of w, and 2.2, in the time-domain give rise to peaks in the frequency-domain, referred to as sidebands, which appear at integral multiples of the spinning speed from a centerband. The intensities of sidebands that are related to the CSA,“‘-15 and enable one to mea- sure the chemical shielding tensors. 4. Methods for Measuring Chemical Shift Anisotropy (CSA) from MAS-NMR Spectra In this study, three methods were used to calculate the principal components of the chemical shielding tensor: the moments method of Marlcq and Waugh,14 a graphical procedure developed by Herzfeld and Berger that is based on spectral simulation,15 and a MAS-NMR spectral simulation program. A. Moments Method The I-th moment of an NMR spectrum“ is obtained using the following definition (integral) and approximation (summation): M, = f _°:°w’g(w)d.a 2 WrIENng(Nwr) "'24) with w the frequency of an isochromat (with respect to the isotropic chemical shift position), g(w) the intensity of the area normalized to one at frequency u), N the 21 order of the N-th sideband (positive or negative integer, 0 for the centerband), g(Nw,) the normalized area (or peak height) of the N-th sideband at frequency N10,; and w, the spinning frequency. For a reasonably “sharp” experimental MAS- NMR spectrum, the summation in Eq. I-24 yields moments very close to the true moments defined by the integral expression The moments method yields the components of the chemical shielding tensor by relating them to the measured second and third moments of the MAS-NMR spectrum with the following equations:“ M2 = (62/15)(3 + 772) (1-25) M3 = (253/35)(1 — 7,2) (I-26) where M2 and M3 are the second and third moments respectively, and 6 and the asymmetry parameter 1; (Eq. l-6) are related to the shielding tensor components. ' 6 and an isotropic chemical shift are represented by the following equations: 5 = 033 - 0,- (1-27) 012%(‘711‘1'0221'030- (1'28) Using experimentally determined values of the second and third moments, and of the isotropic chemical shift (0,), we can determine the individual components of the shielding tensor 011, 022, and 033. The chemical shift anisotropy (CSA) can be calculated from Eq. l-7. B. Herzfeld and Berger Graphical Method The Herzfeld and Berger graphical method was used to obtain the chemical shielding tensors by overlaying the various calculated contours corresponding to the measured ratios lg/lo, where l,, is the intensity of the i-i-th sideband and I0 is the intensity of the centerband. The region of overlap provides both values 22 and error limits for the intermediate parameters 8 and p, which are then used to calculate the chemical shift tensor from equations (I-29) and (I-30):15 it = (7H0) (033 - 0111/90 (1‘29) P=(011 +033-20221/(033-0111- (1'30) 5. Experimental A. MAS-NMR Studies Alkaline earth fluorapatites were synthesized and provided by Dr. Chung- Nin Chau of GTE Chemicals, Towanda, PA. The 19F MAS-NMR spectra were recorded at 376MHz on a 9.4T Varian Associates VXR-400 spectrometer at the Max T. Rogers NMR Facility at Michigan State University. The fluorine radiofre- quency was amplified by an AMT model 3137/3900-2 amplifier. A high speed 19F MAS-NMR probe from Doty Scientific with 5 mm o.d. Si3N4 rotors with Vespel caps was used. The spinning speed was measured with a fiber optic detector, and was constant to within :10 Hz during acquisitions. The magic angle was set by minimizing the linewidth of calcium fluorapatite, which also provided a sec- ondary chemical shift reference (64.0 ppm with respect to hexafluorobenzene at 0.0 ppm).6 The 1r/2 pulse length was 4.0-4.2 83, and spectral widths of 100 - 120 kHz were used. An exponential apodization corresponding to a line-broadening of 37-80Hz was applied to the free-induction decay, which was the result of four scans with a relaxation delay greater than five times the spin-lattice relaxation time T1. The T1 values of centerbands were obtained by using an inversion-recovery sequence, and are 85, 87, 88, 101, and 112 second for Cas F(PO4)3, Sr5F(PO4)3, 385 F(PO4)3, 083973133 F2(PO4)6, and Ca58r5F3(PO4)3 respectively. The SPAR- 23 TAN pulse sequence, shown in Fig. I-4, was employed for 19F MAS-NMR spectral spin diffusion measurement.20 The centerband and assorted sidebands were se- lectively inverted by a 180° DANTE pulse train21 consisting of twelve 15° (2 8s) pulses given at the same point of each rotor cycle (rotor-gated synchronization). DANTEPulses fit/2 lllllll—T’“ acquire Figure I-4. SPARTAN pulse sequence for monitoring spectral spin diffusion.20 (see text) 24 The power levels and pulse lengths of the pulses in the DANTE train were adjusted to make the excitation profile suitably selective. After the mixing period, a nonselective 90° read pulse (12 89) was given with alternated phases to cancel out imperfections in the DANTE pulse trains22 that resulted in incomplete inversion of the magnetization at a specific frequency. The peak intensities used to calculate the moments at the different spinning speeds were obtained both from integration and from deconvolution of the indi- vidual peaks using VNMR 3.2 software, and the peak intensities from integration rather than peak heights were used for the Herzfeld and Berger analysis, since the half-height linewidths (Al/1,2) of the centerband and sidebands can be some- what different. All 19F NMR spectra were baseline-corrected prior to integration and deconvolution to remove a “dip" presumedly due to receiver overload. Since there are impurity peaks that overlap the peaks of strontium fluorapatite and bar- ium fluorapatite in the 19F NMR spectra, the CSA values determined by using the deconvolution data are assumed to be more accurate than those obtained using integrated intensities. A Lorentzian shape was assumed for the deconvolutions, and the half-height linewidth and frequency of each peak was allowed to vary. 8. Simulation of 19F MAS Spectra of Fluorapatite Samples Simulations of the 19F MAS-NMR spectra that take into account of the CSA but not dipolar couplings were performed by using the MAS-NMR simulation routine provided with the Varian VNMR 3.2 software. The input parameters are the chemical shift anisotropy, the asymmetry parameter, the isotropic chemical shift, the Lorentzian linewidth, and the spinning speed. The peak widths of the simulated spectra were chesen to agree with the experimental ones, and 8K complex points and 8K zero-filling were used. Alternatively, the PC-based 25 computer program ANTIOPE,23 which takes account for dipolar couplings, was used to simulate the spectrum of five linear spins 3.44 Angstroms apart, with a CSA tensor corresponding to that of Ca5F(PO4)3. The ANTIOPE simulation was performed by observing magnetization of the middle spin (+3) of a 5 spin system, since the homonuclear dipolar coupling pattern between the middle spin and its neighbors better reflects the couplings of the infinite linear spin chains in Ca5F(PO4)3. The 256 complex data points calculated were zero-filled to 8K to avoid errors from inadequate digital resolution and apodized with a Iinebroadening of 200Hz. 6. Results A. Alkaline Earth Fluorapatites Table 1 shows the measured 19F MAS-NMR spectral moments of Ca5F(PO4)3, Sr5F(PO4)3 and Ba5F(PO4)3, and the calculated CSA and r) values at different spinning speeds obtained from the moments method. The existence of imaginary values of r), obviously lacking any physical significance, arises from experimental errors in the moments measurements. From equations (l-7), (l-27), and (I-28), we can rewrite the CSA in the following form; CSA=%(033 —0,.) :38. (181) Therefore, we can calculate the value of a CSA without knowing the value of n. Theoretically, the second and third moments are independent of the spinning speed.14 The changes in the second and third moments with spinning speed seen in Table 1 are the result of experimental error; therefore, we use the average value of the moments over all the spinning speeds to calculate an average CSA and n. 26 Table 1. 19F NMR chemical shift anisotropy and asymmetry parameter 1) of M5F(PO4)3 (M = Ca, Sr, and Ba) calculated using the moments method, from both integration and deconvolution. The numbers in parentheses are obtained from the integration data. u, M1 M2 M3 Measured Measured (kHz) (ppm) ((ppm)2) ((ppm)3) 7) CSA (ppm) 4.12 -0.25 660 10690 0.092 36-0 (0.20) (684) (1 1740) (0.2191) (374) 5.12 -0.09 662 1 1340 0.167i 86.7 (0.14) (663) (11950) (0.2481) (87.3) 6.12 -0.13 651 10980 0.150 85.9 (0.20) (653) (11830) (0.2661) (86.8) 7.12 -0.08 646 10610 0.0951 85.4 ,2 (-0.12) (649) (10480) (0.073) (85.4) V g 8.12 -0.03 636 10310 0.0681 84.6 ‘3’ (-0.06) (638) (10470) (0.1 10) (84.9) o 9.12 -0.01 638 104W __§—_0.081 W. (0.18) (634) (10800) (0.1951) (85.0) 10.18 0.01 643 10440 0.0081 85.1 (0.10) (643) (10290) (0.062) (84.9) = :l f 3 average 0081 648 1 0680 0.0951 85.5 0.08 1:12 1340 (0.1791) 10.7 (0.0901: (649: (11080 (85.6 0.08) 1 1) i660) 1:1 .0) — 4.29 -0.48 20690 0.204 108.4 (0.05) (1°69) (22530) (0.08i) - (109.8) 5.63 -0.21 921 17700 0.077 101.7 (0.45) (917) (17090) (0.158) (101.2) 27 6.26 -0.37 1054 20630 0.194 108.2 (0.83) (1036) (19460) (0.243) (106.9) 7.00 -0.30 1077 23930 0.2011 1 10.8 (0.52) (1095) (24340) (0.1841) (1 1 1.6) ,3 8.20 -0.38 1036 19940 0.207 107.2 2' (0.54) (1025) (18770) (0.268) (106.1) 11; 10.20 0.11 1033 21180 0.026 107.8 :71 (0.16) (1047) (22110) (0.121 i) (108.2) average -0.26 1030 20690 0.117 107.4 10.21 151 11820 (0.119) 12.8 (0.18 (1032 (20720 (107.5 10.46) 156) :I:2480) 13.3) _ 6.23 0.10 2314 69840 0.110 161.8 (0.74) (2304) (73820) (0.1701) (161.0) g 8.19 0.02 2209 58000 0.267 157.2 g (0.53) (2248) (61030) (0.295) (155.4) “a 10.10 0.03 2227 63410 0.190 157.3 3 (0.90) (2268) (70380) (0.150) (159.0) 1087 0.39 2289 70220 0.050 160.5 (0.71) (2326) (70470) (0.110) (161.5) =# average 0.00 2270 66130 0.170 159.0 10.33 134 14000 (0.1 12) :l:1.8 (0.59 (2277 (68170 (159.4126) 10.34) 145) 16030) # 28 The values for the CSA and moments in parentheses In Tablet are those deter- mined from integration data, and the standard deviations are larger than those obtained from the deconvolution data. The CSA values can be also obtained by using the graphical procedure of Herzfeld-Berger, which involves measurement of the sideband intensities. Fig. l-5 shows the two graphical plots of Herzfeld-Berger for Ca5F(PO4)3 spinning at 6.12 and 10.18 kHz. The measured intensity ratios Iii/Io of Ca5F(PO4)3 spinning at 6.12 kHz do not overlap at any region of the plot. At the higher spinning speed of 10.18kHz, a region of overlap is observed, centered around p = 0.95 i 0.05 and p = 3.19 :l: 0.19. Other plots using the Herzfeld and Berger graphical method for Sr5F(PO4)3 spinning at 6.26 and 10.20 kHz are shown in Fig. l-6. From Fig. l-6a, only four lines out of the ten lines overlap around p = 0.46 i 0.05 and 8 = 6.4 :i: 0.20, and the CSA and 17 value (92 1 4 and 0.47 1 5) are quite different from those obtained at a 10.20 kHz spinning speed (see Table 2). However, the overlap of the contour lines of Iii/l3 for Sr5F(PO4)3 spinning at 10.20 kHz occurs at around p = 0.96 :t 0.04 and p = 4.0 :t 0.19. Fig. l-7 shows the Herzfeld and Berger plots of Ba5F(PO4)3 at two different Spinning speeds (6.23 kHz and 10.18 kHz). Unlikely the other fluorapatite sam- ples, all contours of Ba5F(PO4)3 at 6.23 kHz except for L3 intersect at one point around p = 0.8 i 0.04 and 8 =- 10.4 1: 0.13, and those at a 10.18kHz spinning speed intersect at around p = 0.78 :t 0.02 and 8 = 6.2 1 0.13. The p and 8 values of the intersection points are used to obtain the chemical shielding tensors (CSA and n) from Eq. (l-29) and (I-30). 29 1.0 b) Figure l-5. Contour plots of Ca5F(PO4)3 spinning at 6.12 kHz (a) and 10.18 kHz (b) using the Herzfeld and Berger graphical method. 30 1.0 b) -O.5 ‘ -1.0 _ 1.0 A 0.5 1 P 0.0 ‘ Figure l-6. Contour plots of Sr5F(PO4)3 spinning at 6.26 kHz (a) and 10.20 kHz (0) using the Herzfeld and Berger graphical method. 3O 1.0 b) 0.5 '1 -O.5 ' -1.0, T ' 1.0 a) Figure l-6. Contour plots of Sr5F(PO4)3 spinning at 6.26 kHz (a) and 10.20 kHz (b) using the Herzfeld and Berger graphical method. 31 1.0 b) 0.5 ‘ -1.0 , 1 1.0 0.5 ‘ -0.5 Figure. l-7. Contour plots of Ba5F(PO4)3 spinning at 6.23 kHz (a) and 10.18 kHz (0) using the Herzfeld and Berger graphical method. Table 2. The 19F chemical shift anisotropy (CSA) and asymmetry parameter r; of M5F(PO4)3 obtained using the Herzfeld and Berger graphical method and integrated peak intensities. 32 u,(kHz) Measured 77 Measured CSA (ppm) Ca5F(PO4)3 10.18 0.04 1 0.04 86 1 5 Sr5F(P04)3 10.20 0.06 1: 0.04 107 1 7 6.23 0.16 :l: 0.03 164 1 4 BagF(PO4)3 8.19 0.23 1: 0.04 157 1 7 10.10 0.18:l:0.01 15814 10.87 0.201: 0.04 160 i 5 Table 3. 19F chemical shielding tensors of M5F(PO4)3, calculated from the moments method using peak intensities obtained from deconvolution. An average value of the CSA at various spinning speeds was used, along with an assumed 17 value of zero, to obtain the principal components with respect to C3F3 and free F' ion (parentheses). 011 (ppm) ( = 022) 033(ppm) Ca5F(PO4)3 35.5 1 0.2 121.0 1 0.5 ( 159.5 1 0.2 ) ( 245.0 1 0.5 ) Sr5F(PO4)3 61.4 1 1.1 168.8 1 1.9 (185.4:l:1.I) (292.8119) Ba5F(PO4)3 131.8106 290.8 1 1.2 (255.9106) (414.9112) 33 Table 2 shows the CSA and 17 values obtained by the graphical procedure of Herzfeld-Berger. The standard deviations are determined by the intersection area (box size in Fig. I-5, I-6 and l-7). The contours of the experimental ratios Iii/Io for Ca5F(PO4)3 and Sr5F(PO4)3 fail to overlap at spinning speeds below 9kHz but they overlap at 10.18 and 10.20kHz respectively. Therefore, the CSA values of CasF(PO4)3 and Sr5F(P04)3 are obtained at the high spinning speed. The Iii/Io contours of Ba5F(P04)3 overlap within a small region, and the resultant CSA values are close regardless of the spinning speed. Even though the measured asymmetry parameters of M5F(PO4)3 are different and non-zero at various spinning speeds, the chemical shift tensors of M5 F(PO4)3 must be axially symmetric on the basis of the X-ray crystal structure."'11 Therefore, we constrain r) to be equal to 0 and use the average CSA determined using the moments method to calculate the chemical shift tensors in Table 3. It is necessary to know the chemical shift tensor values on an absolute chemical shielding scale with respect to free fluoride ion in order to be able to calculate the contribution of the sigma- and pi-bonding to the paramagnetic shielding. The isotropic chemical shift of C3F3 has been calculated to be 124 ppm with respect to free fluoride ion.‘9 The chemical shift tensors of M5F(P04)3 with respect to both C3F3 and free F' (parentheses), obtained using the average CSA from the deconvolution data in Table 1 and assuming 1) = 0, are shown in Table 3. Experimental and simulated 19F MAS-NMR spectra of M5 F(PO4)3 samples at spinning speeds near 6kHz, using the chemical shift tensor components in Table 3, are shown in Fig. l-8. The isotropic chemical shifts of M5F(PO4)3 ( M = Ca, Sr, and Ba) are 64.0, 97.2, and 184.8 ppm from hexafluorobenzene respectively. The half-height linewidths of the centerbands of Ca5F(PO4)3, Sr5F(PO4)3, and 34 Ba5F(PO4)3 without line-broadening are 164 Hz, 385 Hz and 438 Hz respectively. Brunner et. al. have used average Hamiltonian theory to provide a general expression for the MAS-NMR linewidths of spins with axially symmetric shielding tensors.24 Their expression takes into account homo- and heteronuclear dipolar interactions and CSA, and predicts a half-height linewidth (HHLW) that is inversely proportional to the spinning speed. Figure l-9 shows a plot of the 19F MAS-NMR centerband HHLW vs. spinning speed for M5F(PO4)3 (M = Ca, Sr, and Ba). The HHLW of Ca5F(PO4)3 decreases monotonically with increasing spinning speed within experimental error, but those of Sr5F(PO4)3 and Ba5F(PO4)3 do not. The dependence of the linewidth of M5F(PO4)3 upon spinning speed indicates that the linewidth is broadened homogeneously. The “plateau” value of the HHLW increases as the CSA increase (from Ca to Sr to Ba); this may simply reflect _ larger effects of crystal imperfections upon the isotr0pic shifts of Sr5F(PO4)3 and Ba5F(PO4)3. 8. Solid Solutions of Ca/Sr Fluorapatite The 19F MAS-NMR spectrum of Ca3,g-,Sr1_o3F2(PO4)3 spinning at 8.23 kHz is shown in Fig. l-10a. It is very difficult to unravel the isotropic chemical shifts and their sidebands at a fixed spinning speed due to the overlap of the different centerbands and sidebands. Since the sidebands are located at Integral multiples of the spinning speed from the centerband, we can differentiate between the centerband of one peak and a sideband from another peak with a different isotropic chemical shift simply by increasing the spinning speed. In this way, three isotropic chemical shifts of the peaks in Ca3_97$r1,33F2(P04)3 (Fig. l-10b) can be obtained, at 64 ppm, 79.6 ppm, and 97 ppm. Overlap of centerband and sideband peaks makes it difficult to obtain reliable integral intensities of individual peaks in the 35 _____._J__.LJ__LJL_JL_.__ c}. 1 LLJLJLL1__ LL_ “1.1111011. a) e ‘— V T V T T f r T r 1 1 r r I Y r r r r f T T r T T *1 v 300 280 200 150 100 50 pp- Figure l-8. The experimental (lower) and simulated (upper) spectra of M5 F(PO4)3. a) BagF(PO4)3 at 6.23kHz; b) Sr5F(PO4)3 at 6.26kHz; c) Ca5F(PO4)3 at 6.12kHz. ' indicates the centerband (0,). Simulation were based on the chemical shift tensor components in Table 3, and neglect dipolar coupling. Hall-linewidth (Hz) 36 600 n 500‘ a :1 1.1 an 1:: 4m'DAAA 1:1 ,0 A D :1 0'3 a 300- °o a 1 A A A A A °o .. o 200 O O 4 o 00 100 1 1 . fir w - r - 2 4 6 8 10 12 Spinning speed (kHz) Figure l-9. ‘9F MAS-NMR center band HHLW vs. spinning speed for M5F(P04)3 (M . Ca, Sr, Ba). CasF(PO4)3 (open circle), Sr5F(PO4)3 (open triangle), and Ba5F(PO4)3 (open square). 37 spectrum. From Fig. I-10b, only the integrated intensity of the centerband and sidebands of the 79.6 ppm peak can be measured since the other centerbands and sidebands overlap. The ratio of the integrated intensity of the centerband and sidebands of the 79.6 ppm peak to the total integral intensity in Fig. l-10b is 29 %. The deconvoluted spectra of the 64 ppm peak and also the peak near 51 ppm due to the -1 sideband of the 79.6 ppm peak of Fig. I-10b are shown in Fig. l-11. Three peaks can be deconvoluted from the asymmetric 64 ppm peak. The gaussian fractions of peaks l and II are 0.88, and that of peak III is 0.68. The half-height linewidths of deconvoluted peaks I, II, and III are 454, 536, and 758 Hz respectively. The percentages of the integral intensity of deconvoluted peaks I, II and III are 45.4, 37.3 and 17.4 % Fig. l-12 shows the 19F MAS-NMR spectrum of Cag_978r1,33F2(PO4)3 obtained using the SPARTAN pulse sequence at 10.87 kHz.20 The transmitter offset of this experiment was set to invert the 64 ppm peak, leaving the other peaks unperturbed. Since the pulses in the DANTE train are repeated every rotor period, the sidebands of the 64 ppm peak are also inverted. The mixing times are 0.1, 1, 3, 6, 9, 12, 15, 18, 21, 24, and 30 seconds. As the mixing time increases, the intensities of adjacent peaks decrease while those of the centerband (and sidebands, not shown) of the 64 ppm peak recover quickly compared to the spin- lattice relaxation time of Cag,gysr1,o3F2(PO4)3 (T 1 = 101 second). Fig. I-13 show the 19F MAS-NMR spectra of Cassr5F2(PO4)3 spinning at 8.25 kHz and 10.30 kHz. The centerbands are resolved by spinning at 10.30 kHz. The values of the isotropic chemical shifts in CaSSr5F2(PO4)3 are 69.7ppm, 86.8 ppm, and 104.5 ppm, which are about 6 ppm downfield compared to those in Ca3,97$r1,o3F2(PO4)3. The measurement of the integral intensity of peaks is 38 m l @008me U Q I I a) i l g \ : T'Tr'V'TrYFTTV'_'TTYTWTTWITTTTTVVf'rv—iriITrrYIYIYTTvvvvyvrrrrvvr'IV'vvi'YY'it‘V'iv'Y' rrVTTT" 180 160 140 120 100 80 60 40 20 0 ppm figure I-10 19F MAS-NMR of Cag,g-,Sr1,o3F2(PO4)3 spinning at 8.23 kHz (3) and at 10.80 kHz (b). ' indicates the centerband. 39 :5 '\/ \ I'T'VITYTrTTTV‘ITT'V r1 VTYYII'II'TT TIT . 11111111'111111 ‘r ITI'TTTTTTTTT‘TT ITITTTI'TTIIIYT r . (11111111111111 [1111] 1111711111 ITTTTITIVVTITIIVV . I'VTTTTT 68 66 64 62 6 0 58 56 54 5 2 50 48 46 44 PP'“ figure I-11. Deconvolution spectrum 0164 ppm peak in fig. I-11b ‘fi—‘lij 60 55 50 ppm [FTTfT 80 75 70 55 Figure l-12. 19F MAS-NMR spin diffusion experiment of Ca3_978r1_o3F2(PO4)6 using SPARTAN pulse sequence spinning at 10.87 kHz. Mixing times are 0.1, 1, 3, 6, 9, 12, 15, 18, 21, 24, and 30 seconds. (899 text) I l'l'l'Il'I‘I'Il l l I l 140 100 30 60 40 20 O -20 ppm , . 180 figure l-13. 19F MAS-NMR of Ca5$r5F2(PO4)3 spinning at 8.25 kHz (3) and at 10.30 kHz (b). ‘ indicates centerband. 42 hindered by the overlap of the centerbands and sidebands. 7. Discussion A. Measurement of the 19F Chemical Shift Anisotropies of M5F(PO4)3 Using Different Methods The values for M5F(PO4)3 are obtained by using two different methods, the moments method and the Herzfeld and Berger graphical method. In the moments method, the accurate measurement of the second and third moments of MAS- NMR spectra is difficult in general because of the low signalrnoise ratio of the weak higher order sidebands that contribute significantly to the moments.25 The homo- and heteronuclear dipolar interactions also contribute to the experimental second moment, making it difficult to separate their contribution to the second moment from that of the CSA alone. We will discuss how the dipolar interaction influences the measurement of the CSA using the moments analysis method. MAS-NMR simulations using ANTIOPE (Fig. l-14a) show that the including homonuclear dipolar interactions result in a higher intensity of the centerband relative to that of the sidebands. The contribution of the homo- and heteronuclear dipolar interactions and of the CSA to the second moment is not simply additive.“ The values of the CSA obtained from the simulated 19F MAS-NMR spectra calculated using the CSA alone (VNMR 3.2) and using the CSA along with dipolar interactions (ANTIOPE) differ by approximately 2 ppm at the spinning speed of 6.12 kHz. Since the low signalznoise ratio of the weak higher order sidebands gives rise to measurement errors in the experimental and since considering the dipolar interaction can slightly change the values of the CSA, we use the moments measured from the intensities of the sidebands of the experimental spectra to 43 obtain the values of the CSA in Table 1. When we use the Herzfeld and Berger graphical method, the chemical shift tensors are obtained from the coordinates of the overlapping contours of the values of I,,,/lo ratios. Intensities of all peaks (a centerband and sidebands) in a MAS spectrum are used to obtain a CSA with the moments method, whereas those of a limited number of peaks are used to obtain a CSA with the Herzfeld and Berger graphical method (lg/lo ratios for up to :I: 5 sidebands).15 Thus, the measurement error of the CSA obtained from the Herzfeld and Berger graphical method should beless severe than that from the moments method owing to the neglect of weak higher order sidebands. The lfi/Io ratios of the samples with a larger CSA and small dipolar interactions have a smaller relative contribution from the dipolar interactions. The dipolar interaction decreases on going from Ca5F(PO4-)3 to Ba5F(PO4)3 as a result of the expansion of the lattice,911 whereas the CSA values increase going from Ca5F(PO4)3 to Ba5F(PO4)3. Therefore, it is not surprising that the contours of lfi/lo for Ba5F(PO4)3 overlap at one point or over a small region at all spinning speeds. Likewise, the contours of l,,,i/lo for Ca5F(PO4)3 and Sr5F(PO4)3 fail to overlap below about 9kHz due to the greater size ratio between the homonuclear dipolar interaction and the CSA. As the spinning speed increases, the higher order sideband intensity contribution from the dipolar interaction appear to decrease quickly. Thus, most of the intensity of the higher order sidebands at high spinning speeds is due to the CSA. Figures I-5 and 6 show that the value of 8 of the +1 sideband is insensitive to a change of the spinning speed. The values of 8 of the other sidebands move to the value of 8 of the +1 sideband with an increasing in the spinning speed. The integrated intensity of each peak in the experimental spectrum does not 44 correspond to the peak height because of differences in the half-height linewidth of the peaks. For example, the peak height of the centerband in Gas F(PO4)3 is larger than that of the -1 sideband; however, the integrated intensity of the centerband in Ca5F(PO4)3 is less than that of the -1 sideband since the half-height linewidth of the centerband (164 Hz) is smaller than that of the -1 sideband (217Hz) (Fig. MO). The ratio of the integrated intensity of the centerband to that of the -1 sideband is 1.00:1.13 in Fig. l-10. Thus, when CSA is obtained by using MAS- NMR, the integrated peak intensity must be used rather than a peak height to obtain a more accurate CSA value. Since the linewidth of the sidebands in the experimental spectra vary for a given sample, it is convenient to represent these spectra as stick spectra for comparison with simulated spectra Stick spectra (Figure I-14) normalized to the -1 sideband of M5F(PO4)3 were obtained from simulation (ANTIOPE and VNMR 3.2) with the parameters of Table 3. The intensities of the centerband and sideband of the simulated stick spectra taking into account only chemical shift anisotropy (VNMR simulation) are different from those of the experimental spectrum in Fig. I-14a. The simulated spectrum obtained using ANTIOPE (CSA and dipolar interaction among uniformly-spaced five fluorine spins) more closely resemble the experimental spectrum in Fig. l-14a. Fig. l-14 shows that the intensity differences between the centerband and sidebands of the simulated and experimental stick spectra become smaller and smaller less going on from Ca5F(PO4)3 to Ba5F(PO4)3. B. Separation of Metal-Fluoride Sigma- and Pi-Bonding Contributions to the 19F Shielding Tensor The chemical shielding of a nucleus can be separated according to Ramsey’s 45 figure I-14. Comparison of simulated and experimental 19F MAS-NMR stick spectra of M5F(PO4)3 at the same spinning speed. The asymmetry parameter 1) is forced to be zero and the average CSA values of M5F(PO4)3 obtained from the moments method are used, and all peaks are normalized to the -1 sideband. 46 1.2 '2 . 31.0 E/ 9 1 =/ .6 E/ 08- =/ 1'; 7‘/ /-/ 9 1 8_é N - =/—/ -.-_- =¢—/ to 1 E%—é E 04- =/ E/_¢ 3 ° =/ E?—/ . E/ =/—¢ '2 «E; -/E Eé—/=/ 5 0.21 _é§ Eé=égé '0.0' éE E%_%Eé Order of sidebands CI The intensity (from deconvolution) of the centerband and sidebands of the experimental spectrum. E The intensity of the centerband and sidebands due to the CSA only (VNMR 3.2 simulation). The intensity of the centerband and sidebands due to the CSA and dipolar interactions among 5 fluorine spins (ANTIOPE simulation). 3) Simulated and experimental 19F stick spectra of CasF(PO4)3 Spinning at 6.12 kHz 47 1.2 E 10 3 . Q .9 f.’ 0.81 .0 E 0.6- '5 ‘ = E 0.4- E 5 ‘ E g 0.2- E]: s I 1.11 =11 s1 .. 0.0 , , es ,, 3 2 10-1-2-3-4 -5 Order of sidebands CI The intensity (from deconvolution) of the centerband and sidebands of the experimental spectrum. E The intensity of the centerband and sidebands due to the CSA only (VNMR 3.2 simulation). b) Simulated and experimental 19F stick spectra of Sr5F(PO4)3 spinning at 8.20 kHz 2.0 1.0 'l 1 0.0 mafia-IE] EL“ 876543210-1-2-3-4-5-6 Intensity normalized to -1 sideband j lllIIIllllllllllllllllllllllllI[IIIIIIIIIIIIII llllll IllIIllllllllllllllllllllIIII Ill[lllllllllllllllllllllllll I llllllllllllllll I El f T Order of sidebands CI The intensity (from deconvolution) of the centerband and sidebands of the experimental spectrum. E The intensity of the centerband and sidebands due to the CSA only (VNMR 3.2 simulation). c) Simulated and experimental 19F stick spectra of Ba5F(PO4)3 spinning at 8.19kHz 49 formulation“27 into a diamagnetic term and a paramagnetic term. Lamb showed that the diamagnetic contribution to shielding is proportional to the sum of in- verse distances between the i-th electrons and the nucleus.28 The calculated difference in the diamagnetic term for free fluoride ions in different ionic fluo- rides is small since the distances of the i-th electrons to the nucleus for a free F' ion are similar to those for different ionic fluorides.27 Therefore, differences in the paramagnetic term are largely responsible for the fluorine chemical shift changes observed in various ionic fluorides. The chemical shielding tensors of the paramagnetic term, measured with respect to free fluoride ion are used to separate metal-fluoride sigma— and pi-bonding contributions to the 19F shielding. The paramagnetic contribution depends upon the electronic ground and excited states. The paramagnetic contribution of electrons in s orbitals can be ignored, since the angular momentum of s orbitals is zero. An asymmetric distribution of p and d electrons near the nucleus, and low-lying excited states of these electrons, can result in a large paramagnetic term.26'27 A superposition of the ground and excited states of sigma-bonding orbitals arises when an external magnetic field is applied. The angle between a single sigma-bond and the extemal magnetic field determines the mixing of the states. No perturbation of the symmetry of the electron cloud occurs when the external magnetic field is along the sigma-bond. The shielding constant arising from the paramagnetic term in linear compounds depends on the angle between the external magnetic field and the direction of the sigma-bond in the following way.25'29 0(6) = aasinzél (I-32) where 0(6) is the paramagnetic shielding constant at angle 0 and a0 represents the contribution of the a-bond to the paramagnetic shielding. Since the pi orbital 50 is perpendicular to the bond direction, the shielding constant due to the pi-bond is represented by 0(6) = awcoszo (l-33) where a“ represents the contribution Of the pi-bond to the paramagnetic shielding, and 0 is the angle between the bond and the external magnetic field. The paramagnetic shielding term arising from both a sigma-bond and a pi-bond is the sum of equations (I-32) and (l-33) 0(0) = aasinze + awcosze = (2/3)aa+ (1/3)a7r+ (1/3)(a7r - ao)(3cos29-1). (l-34) The isotropic portion of the shielding 0,-(9 = 54.7°) is represented by (2/3)aa + (1/3)a7r. The total paramagnetic shielding a is assumed to be the sum of the contributions from the individual bonds; 0 (a): 2 0,-(éj) (I-35) ' Gagarinski els?9 separated the paramagnetic shielding components (a0 and a) of CasF(PO4)3 by using equations l—34 and l-35, and experimental values for the CSA and the isotropic chemical shift. When 033 is parallel to the external magnetic field, the angle between the external magnetic field and the three Ca - F bonds is perpendicular. From the equations I-34 and I-35, 033 (6 = 90°) and (Ii (0 = 54.7°) can be represented by the following equations 033 = 3 [ 2/3 a, +1/3 a1r +1/3(a7r - a0) (3 cos290° - 1) 1 = sag (l-36); ai = Zaa + a3. (I-37) Table 4 shows the ac and a; values calculated with respect to free F‘ ion at zero ppm (since the latter has no paramagnetic contribution to its shielding), whose shift has been estimated theoretically in Ref. 29. We have assumed that the Sr“ ions in Sr5F(PO4)3 form a plane containing the F. ion, since this is known 51 Table 4.19F paramagnetic shielding parameters a; and an for M5F(PO4)3 calcu- lated from the chemical shielding tensors (absolute chemical shift scale) in Table 3. 30(ppm) a1r(ppm) Ca5F(PO4)3 81.7 i 0.1 24.6 :1: 0.3 Sr5F(PO4)3 97.6 i 0.6 26.0 :t 1.2 Ba5F(PO4)3 138.4 :1: 0.4 32.1 i 0.8 52 140 " ‘ 385F(P04)a 100 '4 . S'sF(PO4)3 E Ca5F(PO.)3 so 1 « a. ____F a 20 P - 7" - r - . 1 180 220 260 300 01039“) Figure l-15. a0 and a, for M5F(PO4)3 vs. ‘9F isotropic chemical shift with respect to free F“ ion. The solid lines simply connect the data points; open circles and open smares denote a0 and a“ respectively. 53 to be the case for Ca and Ba fluorapatitef’v11 The increase in ad and an from Ca5F(PO4)3 to Ba5F(PO4)3 indicates more contributions to the paramagnetic shielding, but not “increased bonding” necessary. The a0— and aw values for Ca5F(PO4)3, Sr5F(PO4)3, and Ba5F(PO4)3 are plotted vs. the isotropic chemical shift with respect to free F' ion in Fig. l-15. Fig. I-15 shows that the slope of a0 vs. the isotropic chemical shift is larger than that of a3 vs. the isotropic chemical shift. This observation implies that the observed increase in the paramagnetic shielding term as one goes from Ca to Sr to Ba in M5F(PO4)3 (M = Ca, Sr, Ba) is due to primarily to an increase in the sigma-bonding parameter a0. C. Study of Site-Preference of Sr2+ Ions in Ca/Sr Fluorapatite Solid Solutions Using 19F MAS-NMR The difference in the chemical bonds between a fluoride ion and either calcium or strontium gives rise to the different 19F chemical shifts of fluoride ions. The flu- oride ions of Ca3.37$r1,o3F2(P04)3 have the four different chemical environments shown in Fig. l-16. Since the structures of A and D represent the local fluoride ion environment in Ca5F(PO4)3 and Sr5F(PO4)3,9=‘° respectively, we assign the isotropic chemical shifts of A and D in Fig. 3 to 64 and 97 ppm respectively. The isotropic chemical shifts of B and C in Fig. l-16 are predicted using Eqs. l-34 and I-35, and are equal to the sum of the isotropic portions of the shielding of the appropriate number of Ca - F and Sr — F bonds. The configurations B and C in Fug. ~l-16 have two Ca - F bonds and one Sr - F bond, and one Ca - F bonds and two Sr - F bond respectively, resulting in predicted isotropic chemical shifts of B and C in Fig. l-16 of 75.1 ppm and 86.1 ppm respectively. The isotropic chemical shift of either one is different from the isotropic chemical shift of the peak at 79.6 ppm in Fig. l-10b. There are no centerbands between the centerband at 64 ppm 54 Ca Sr Ca Ca Ca Ca A a Sr Sr Sr Ca Sr Sr c D Figure l-16. Substitution of Sr2+ in the three nearest neighbor Ca(2) sites of Ca/Sr fluorapatite solid solutions. 55 and the centerband at 79.6 ppm, and the isotropic chemical shift at 79.6 ppm is the second isotropic chemical shift in the downfield direction in Fig. l-10b. Therefore, we assign the isotropic chemical shift of the configuration B (Sr1Ca2F) in Fig. MS to the 79.6 ppm peak in Fig. I-11b. The fact that the experimental shift of the configuration B is 4.5 ppm downfield of predicted value is reasonable, since the Sr3F configuration in the Ca58r5F2(PO4)6 solids solution is also downfield (by 7ppm) of the shift in Sr5F(PO4)3, presumably due to lattice distortion effect. The isotropic chemical shift due to C in Fig. MS may be between 79.6 ppm and 97 ppm, and may be concealed by overlap with the downfield sideband of the 64 ppm peak in Fig. l-10b. The degree of possible preference of Sr2+ ions for a Ca(2) site for Ca/Sr fluorapatite solid solutions can be studied by using both the resolved peaks obtained from 19F MAS-NMR and from peak deconvolution of overlapping peaks. There are two different types of calcium ions in Ca1oF2(PO4)3. Since there are 4 Ca(1) ions and 6 Ca(2) ions in two unit cells, we rewrite the formula of Ca3,978r1,o3F2(P04)6 as Ca(2)6xSr(2)6yCa(1)4X,Sr(1)4), F2(PO4)6( 6x + 4x’ = 8.97 and 6y + 4y’ = 1.03, and x + y = x’ + y’ = 1). Since the fluorine ions are bonded to Ca(2) and Sr(2) ions, the chemical shift of the fluorine depends on the numbers of Ca(2) and Sr(2) ions to which it is bonded. The integrated areas of the peaks at each chemical shift in Fig. l-16 are represented by the following equations A = x3, B = 3x2y, (l-38) C = 3xy2, D=y3 56 where(A+B+C+D)=(x+y)3=1.0. If the strontium ions randomly substitute for the calcium ions in Ca3.378r1_33F3(PO4)3, the calculated integrated intensity of B from Eq. (l-40) would be 24 %. The ex- perimental value is 29 %, which implies values for x and y of 0.873 and 0.127 respectively. Thus, the experimental integrated intensity of the B peak indicates a 23 % [(0.127 - 0.103)/0.103] site preference of a Sr2+ ion for the Ca(2) site in Ca3,97$r1,o3F2(PO4)3. Table 5 shows the calculated probabilities of the various configurations for both random substitution and a 23% preference of the Sr2+ ions for the Ca(2) site using Eq. (l-38). The 19F MAS-NMR peaks in Fig. l-11b are asymmetric and broad, presumably due to perturbations from strontium ions which are substituted in Ca(1) site, or from strontium ions in the next-nearest Ca(2) sites. We now consider what types of strontium ions (strontium ions in Ca(1) or Ca(2) sites) mainly perturb the fluorine peak at 64 ppm. From the crystal structure of calcium fluorapatite, when a single Sr2+ ion substitutes in a Ca(1) site, it potentially perturbs 6 fluorine ions in an equivalent fashion, but a Sr“ ion substituting for a Ca(2) site perturbs 2 fluorine ions. Thus, the Sr2+ ions in Ca(1) sites perturbs a fluorine ions more than that in Ca(2) sites. The perturbation of the chemical shift of the 64 ppm peak should depend on the number of the strontium ions at the next-nearest Ca(2) sites. The schematic arrangements of the metal cations neighboring a fluorine atom resonating near 64 ppm are shown in Fig. l-17. Since the probability of substitution of more than 2 Sr2+ ions in Ca(2) sites in Ca3,978r1,o3F2(PO4)6 is low, configuration III in Eq. (l-39) represents essentially the sum of the probabilities of having 2 2 Sr2+ ions in Ca(2) sites. The integrated intensities of peaks arising from the individual configurations in Fig. I-17 are then given by the following equations 57 C8 C8 C8 F F /F ' Ca CL .4 C1 I’ Cl C. Ca / F\ F / F C‘ .4 C‘ 9‘ Ca Ca C! 0 Ca / F\ F / ,- Ca .4 Ct ‘ C3 c3 3 .r Figure l-17. Possible configurations for single (l) and double (l) Sr“ substitution in the six next-nearest neighbor Ca(2) sites of Ca/Sr solid solutions of fluorapatite. The middle equilateral triangle in each configuration represents the observed Ca3F group resonating around 64 ppm, and the triangles above and below it (along ' the c-axis), although parallel, are drawn tilted for clarity. One Sr“ ion in the. configuration If is either in a top triangle (as shown) or in a bottom triangle. Two 5,2. ions are either in a top triangle or a bottom triangle, or in both triangles (as shown). 58 Table 5. Calculated probabilities of configurations AD in Ca3_37$r1_03F2(PO4)3 (Figure I-16) for both random substitution and a 23% preference of Sr2++ ions for the Ca(2) sites, and comparison with measured integrated intensity of peak B. Configuration Probability, Probability, Integrated Random 23% Preference MAS-NMR Peak substitution Ca(2) Site Intensity A. 72.1 % 66.5% — B. 24.9% 29.0% 29% C. 2.9% 4.2% — D. 0.1% 0.2% — Total 1 00.0% 99.9% 1 00% Table 6. Calculated probabilities of configurations H" in Ca3.37$r1.33F2(PO4)3 (Figure H?) for both random substitution and a 23% preference of Sr2++ ions for the Ca(2) sites, and comparison with the experimental deconvolution data obtained from the peak near 64 ppm. Configuration Probability, Probability, Intensities of Random 23% Preference Deconvoluted substitution Ca(2) Site Peak at 64 ppm I. 52.1 % 45.6% 45.4% II. 35.9% 38.3% 37.3% III. 12.0% 16.1% 17.4% Total 100.0% 100.0% 100.1% 59 l = x6 II = 6x5y (l-39) Ill = 1— l - II. The comparison of the deconvoluted integrated intensities of the peaks near 64 ppm and the integrated intensities calculated for both random substitutions of Sr2+ ions and for a 23 % preference of strontium ions forth Ca(2) site is shown in Table 6. The predictions for a 23% site preference (x=0.877 and y=0.123) are closer to the deconvolution data than those assuming random substitution. The half-height linewidths of the deconvoluted peaks are broader than those of calcium and strontium fluorapatite (130 Hz at 10.56 kHz, 365 Hz at 10.20 kHz respectively). Since the perturbation effect of strontium ions substituted in the Ca(1) sites on the peak near 64 ppm is small compared to that of strontium ions substituted in the next-nearest Ca(2) sites on the same peak, we believe that the main effect of strontium ions substituting in the Ca(1) site is a slight increase in the half-linewidths of deconvoluted peaks. To prove that the fluorine spins of the peaks at 64 and 79.6 ppm (see Fig. I-10) are actually in the same phase and not in phase-segregated regions, a spin diffusion experiment was carried out. Spin diffusion is the transfer of Zeeman magnetization between two adjacent spins by means of a spin flip-flap?“31 Spin diffusion requires both the existence of a dipole-dipole coupling between the nuclear spins and the conservation of Zeeman energy. Spectral spin diffusion between the peaks having different isotropic chemical shifts under magic-angle- spinning can occur if the Zeeman energy level overlap during a rotor period, the so—called “level-crossing”. The existence of level-crossing between two peaks with different isotropic shifts can be demonstrated by calculating the instantaneous 60 frequencies of the pair of spins in a given crystallite during each part of a rotor cycle.” A MAS Hamiltonian can be transformed from the chemical shift principal axis system (PAS) of each crystallite to a reference frame fixed on the rotor.32'33 The MAS Hamiltonian for the chemical shift principal axis system becomes periodic (cosine wave), and the offset of the cosine wave depends on the asymmetry parameter and the crystallite orientation. Not only is the amplitude of the cosine wave proportional to the CSA but it also depends on the asymmetry parameter and crystallite orientation.”33 Fig. l-12 shows the existence of spin diffusion between the peak at 64 ppm and the peak at 79 ppm. The CSA and asymmetry parameter of the peak at 64 ppm are known.12 Thus, we have to know the CSA and asymmetry parameter of the peak at 79 ppm in order to predict the level-crossing of the two peaks theoretically. We assume that the configuration ll of Fig. I-16 is an equilateral triangle and that the fluoride ion in the configuration If is in the middle of the equilateral triangle. The largest chemical shielding tensor 033 is obtained when the external magnetic field is perpendicular to the equilateral triangle in the configuration ll. Since the shielding principal values 011, 022, 033 are orthogonal, 011 and 022 are on the plane of the equilateral triangle. The smallest chemical shielding component 011 is obtained when the external magnetic field is parallel to a Sr - F bond. From Eq. l-34 and 35, the chemical shielding tensor components 011, 022, 033 of configuration II in Fig. I-16 can be represented by 011 = 0(0°) (Sr) + 0(120°) (Ca) + 0(120°) (Ca) = 3/2 aa (Ca) + 1/2 an (Ca) + a3 (Sr) (|-40) 022 = 0(90°) (Sr) + 0(30°) (Ca) + 0(30°) (Ca) 61 = 1/2 a; (Ca) + 3/2 a (Ca) + a0 (Sr) (l-41) 033 = 0(90°) (Ca) + 0(90°) (Ca) + 0(90°) (Sr) = 2 a0 (Ca) + a; (Sr). (I-42) The values of the shielding tensor componentS011, 022, 033 for the peak at 79 ppm (configuration II in Fig. l-16) obtained using the ad and an values in Table 4 are equal to 160.9 ppm, 175.4 ppm and 261.0 ppm respectively. The CSA and asymmetry parameter of the 79.6 ppm peak calculated using Eqs. l-8 and l-9 are 92.9 ppm and 0.23, and the CSA and asymmetry parameter of the 64 ppm peak (calcium fluorapatite) are 84 ppm and 0.12 The calculated energy modulations of the 64 ppm and 79.6 ppm peaks during one rotor cycle spinning for two different crystallite orientations are shown in Fig. I-18. Fig. l-18a shows the existence of an overlap of the energy levels between the peaks corresponding to configurations l and II in Fig. l-16 during one rotor cycle, whereas the different crystallite orientations of Fig. l-18b do not. From our calculations (not shown here), most crystallite orientations of configurations l and II have an energy overlap during one rotor cycle. Therefore, a powder sample of (Ca3,3-,Sr1,o3F2(PO4)3) will have most crystallites experiencing level-crossing under MAS, resulting in the observed spin diffusion between the 79.6 ppm and 64 ppm peaks. This spin diffusion takes place through the 19F homonuclear dipolar interaction. This spin diffusion between the 79.6 ppm and 64 ppm peaks in Fig. l-12 indicates that the corresponding fluoride ions are close to each other and have equivalent chemical shifts during the rotor cycle. 62 80 60 b) 40 A AAA AAA A A A A A 20 . 0A Ao°°°oA Ao°°°oA O A A O O A A O O A o o o o . oAA AA:3 0AA AAo 0A .20 c,oAMoo 004531300 o E . 063 0C9 0. e '40 t v v E 3° 8. an. 5 A A A 60 1 A a) A A 4 A A A 40 ‘ A A A ‘ A A A 20 -b 000 <90 00 A 00 c’o A oo 00 f o A o o A o o 0 o A o o A o o o A o o -20 °009 9 -40 - . f - 1 - f r 0 0.2 0.4 0.6 0.8 1.0 1.2 Rotor cycle Figure l-18. Variation in the instantaneous chemical shifts during a rotor cycle of the 64 ppm and 79 ppm peaks of the Caa,978r1,33F2(PO4)3 solid solution under MAS for at the different crystallite orientations, a = 0° and fl = 60°(a), and a = 0° and fl - 90° (b): 64 ppm peak (open triangles) and 79 ppm peak (open circles). a and 3 are Euler angles with respect to the principal axis system of tensors in the rotor frame. 63 8. Conclusions We have obtained the values of the CSA for M5F(PO4)3 (M = Ca, Sr, and Ba) using the different methods. Even though the measured asymmetry parameter 77 of Ca5F(PO4)3 obtained by using the moments method and the Herzfeld and Berger graphical method is not equal to zero, the value of the CSA of Ca5F(PO4)3 obtained from 19F MAS-NMR spectra by the two methods is close to that obtained from a 19F single crystal NMR study.12 The accuracy of measuring the CSA of M5F(PO4)3 from the moments remains the same for different spinning speeds. The simulations using ANTIOPE show that the change of the CSA of Ca5F(PO4)3 due to the dipolar interaction is small. The change in the intensity of the center- band and sidebands when the dipolar interaction is considered causes a failure of the Herzfeld-Berger contour plots of Ca5F(PO4)3 and Sr5F(PO4)3 to overlap at spinning speed below 9kHz. However, since the CSA values obtained from the contour plots using high spinning speed data are close to the CSA values measured from the moments method, the two methods are complementary under these conditions. There are two reasons that the intensities of the simulated spectra of M5F(PO4)3 (M a Ca, Sr, and Ba) do not correspond to those of the experimental spectra One reason is that dipolar interaction is neglected in the VNMR 3.2 sim- ulatiOns. The other reason is that the different half-height linewidths of the peaks can give rise to a difference between real and simulated spectra. The stick spec- tra obtained from the integrated intensities of peaks are therefore more useful for comparing the simulated and experimental spectra. The separation of the ad and a3 parameters for from the chemical shift tensors gives information about the contribution of the sigma- and pi-bonding components 54 of the chemical bonding between alkaline-earth metal ions and fluoride ions to the 19F shielding in apatite samples. The values of a0 and a7; obtained from 19F shielding tensors for apatite samples make it possible to predict the CSA and asymmetry parameter r; of configurations in the solid solutions of Ca/Sr fluorapatite and the orientation of the shielding tensor in the molecular frame. This somewhat novel approach to predicting the full shielding tensor in the molecular frame of solid solution has proven valuable in studies of chemical-shift-selective MQ-NMR,34 and may be useful in studying other solid solutions. The shifts of the peaks in the 19F MAS-NMR spectrum of the solid solution Ca8.978r1.o3 F3(PO4)5 were interpreted in terms of proximity of the fluorine spins to strontium ions. The fact that the peaks of this spectrum arose from fluorine spins in the same phase was established by demonstrating the existence of spin diffusion . between them with the SPARTAN pulse sequence. The preference of Sr2+ ions for the Ca(2) sites in Ca3,9-,Sr1,o3F2(PO4)6 was also studied. The observed 23 % preference of Sr2+ ions for Ca(2) sites, using two different methods is in agreement with an X-ray powder diffraction study of Ca/Sr hydroxyapatites, which determined an approximately a 20 % site preference for Ca(2) site.35 A more recent EXAFS study of Ca/Sr hydroxyapatite claimed a larger preference for the Ca(2) site, and must be considered suspect. 65 9. References u—b . M. Mehring, A. Pine, W-K. Rhim, and J. S. Waugh, J. Chem. Phys. 54, 3239 (1971). N . R. W. Vaughan, D. D. Elleman, W-K. Rhim, and L. M. Stacey, J. Chem. Phys. 57, 5383 (1972). 3. R. J. Sears, J. Chem. Phys. 59, 5213 (1973). 4. R. J. Sears, J. Chem. Phys. 61, 4368 (1974). 5. J. Mason, J. C. S. Dalton, 1426 (1975). 6. J. P. Yesinowski and M. J. Mobley, J. Am. Chem. Soc. 105, 6191 (1983). 7. A. T. Kreinbrink., C. D. Sazavsky, J. W. Pyrz, D. G. A. Nelson, and R. S. Honkonen, J. Magn. Reson. 88, 267 (1990). 8. K. A. Smith and D. P. Burum, J. Magn. Reson. 84, 85 (1989) 9. K. Sudarsanan, P. E. Mackie, and R. A. Young, Res. Bull. 7, 1331 (1972) 10. K. Sudarsanan and R. A. Young, Acta Cryst. 828, 3668 (1972) 11. M. Mathew, I. Mayer, B. Dickens, and L. W. Schroeder, J. Solid State Chem. 23,79 (1979). 12. J. L. Carolan, Chem. Phys. Left. 12, 389 (1971). 13. AM. Vakhrameev, S. P. Gabuda, and R. G. Knubovet, J. Struct. Chem. (USSR),19, 256 (1978). 14. M. M. Maricq and J. S. Waugh, J. Chem. Phys. 70(7), 3300 (1979). 15. J. Herzfeld and A. E. Berger, J. Chem. Phys. 73(12), 6021 (1980). 16. R. Tycko, G. Dabbagh, S. R. Kurtz, and J. P. Goral, Phys. Rev. B. 45, 13452 (1991). 17. E. R. Andrew, A. Bradbury, and R. G. Eades, Nature (London), 183 1802 (1959). 66 18. I. J. Lowe, Phys. Rev. Lett. 2, 285 (1959). 19. E. R. Andrew, Prog. Nucl. Magn. Reson. Spectrosc. 8, 1 (1972) 20. L. B. Moran, J. K. Berkowitz, and J. P. Yesinowski, Phys. Rev. B45, 5347 (1992) 21. G. Bodenhausen, R. Freeman, and G. A. Morrison, J. Magn. Reson. 23, 171 (1976); G. A. Morrison and R. Freeman, J. Magn. Reson. 29, 433 (1978). 22. A. Kubo and Charles A. McDowell, J. Chem. Soc. Faraday. Trans. 84, 3713 (1988). 23. F. S. de Bouregas and J. S. Waugh, J. Magn. Reson. 96, 280 (1992). 24. E. Brunner, D. Freude, B. C. Gerstein, and H. Pfeifer, J. Magn. Reson. 90, 90(1990). 25. N. J. Clayden, C. M. Dobson, L. Lian, and D. J. Smith, J. Magn. Reson. 69, 476 (1986). 26. N. F. Ramsey, Phys. Rev. 78, 699 (1950) 27. N. F. Ramsey, Phys. Rev. 86, 243 (1951) 28. W. E. Lamb, Phys. Rev, 60, 817 (1941). 29. Y. V. Gagarinskii and S. P. Gabuda, translated from Zhumal Struktumoi Khimii, 11(5), 955 (1970). 30. N. Bloembergen, Physica15, 386 (1949). 31. A Abragam, The Principles of Nuclear Magnetism (Oxford University Press, London, 1961), Chaps. V. and IX. 32. ET. Olejniczak, S.Vega, and R. G. Griffin, J. Chem. Phys. 81, 4804 (1984). 33. DP. Raleigh, E.T. Olejniczak, and R. G. Griffin, J. Chem. Phys. 89, 1333 (1988). 34. L. B. Moran and J. P. Yesinowski, to be published. 67 35. H. J. M. Heijigers, F. C. M. Driessen, and R. M. H. Verbeeck, Calcif Tissue Int. 29, 127 (1979). PART 2 1H and 19F Multiple-Quantum NMR Dynamics of Quasi-One-Dlmenslonal Spin Distributions in Apatites 68 69 1. Introduction Multiple-quantum NMR is a general term for experiments that observe nuclear magnetic transitions forbidden by the standard selection rule Am = if. Although multiple quantum transitions can be observed by using high power continuous- wave (CW) spectrometers,"14 difficulties of interpretation and inconvenient instru- mental requirements have limited the use of CW-observation of multiple-quantum transitions. The advent of time-domain Fourier transform techniques15 made it possible to detect the forbidden transitions. In the mid 1970s, Hashi16 and Ernst17 independently adapted a two-dimensional Fourier transform NMR technique to in- directly observe multiple quantum transitions. Since that time, the multiple quantum NMR technique has been mostly applied to the liquid state, where the size of the spin systems is relatively small.“3'28 An 8- . pulse sequence that creates an average-Hamiltonian operator for double quantum NMR transitions of dipolar-coupled spin pairs, 29 combined with the use of a time- reversal pulse sequence,30 increases the S:N ratio of multiple-quantum intensities and has made the study of larger spin systems such as found in solids feasible. 1H MQ-NMR has been applied to the study of proton distributions in solids,3‘-32'36 imaging in solids,32"37 and adsorption of organic molecules in zeolites38'39. 1H MQ-NMR has been also used to study nematic liquid crystalsgai‘M3 19F MQ- NMR has been used to investigate fluorine distributions in polycrystalline“ and photosensitive salts.45 The time-development of multiple-quantum coherences in the "infinite" dipolar- coupled spin systems typical of many solids presents both a lure and a challenge. The lure is the possibility of obtaining structural information about groups of cor- related spins that would be otherwise unobtainable with conventional NMR spec- 70 troscopic techniques. The challenge resides in the development of theoretical models describing multiple-quantum dynamics that are both computationally prac- tical and experimentally realistic. The explicit calculation of the density-operator for spin systems, widely used to describe modern NMR experiments,“"5"'8 can- not describe many actual spin systems due to computational limitations (up to nine spins is the current Iimit)."’9‘51 The actual spin systems present in strongly dipolar-coupled solids consist of about 102° spins whose calculation time using the density matrix is prohibitive. Therefore, simplifying approaches have been de- veloped that make only statistical assumptions about the time-development of the density operator, and neglect the detailed spin-dynamics arising from the specific disposition of spins in space.“53 The earliest of these, the statistical model,54 counts the combinatorial possibilities of having coherences of order n ("n-quantum coherences") in a spin system of ”effective size" N. For an "infinite” Spin system, the effective size N increases monotonically with increasing prepa- ration time allowed for creation of multiple-quantum coherences. The statistical model predicts an approximately Gaussian distribution of intensities for the var- ious ordersof multiple-quantum coherence. Although it tends to underestimate the intensity of high order coherences, it does provide a measure of the effective size of the spin system at a given preparation time.3"52'53v55-56 Since intermolec- ular dipolar couplings of nematic liquid crystals are averaged to zero whereas intramolecular dipolar couplings are not, “spin counting” by means of MQ—NMR experiments of the known number of spins in molecules of nematic liquid crys- tals demonstrates the usefulness of the statistical model.“55 The directed walk through Liouville space (“hopping”) model57 predicts the intensities of various or- ders of higher order coherences, at different preparation times, by counting the 71 combinatorial possibilities of allowed "transitions” or "hops” in Liouville space, and assuming that any oscillatory behavior will be hidden by destructive interference and resultant decay. Although the hopping model can successfully account for features of the experimental data,57 Lacelle53 has pointed out aspects of the av- eraging of multiplicative processes that, if properly taken into account, may lead to significant differences from the predictions of the hopping model. Only very recently has there been an attempt to develop a simplifying model, as opposed to explicit density operator calculations?“51 that specifically considers the spa- tial structure of the spin system (including its dimensionality). This "incremental shell" model of Levy and Gleason36 describes multiple quantum dynamics dur- ing the preparation period as a stepwise process: a given coherence can either expand by incorporating one additional spin at the periphery of the spin cluster involved in the coherence, or decrease in size by one spin. The rate of this process is governed by the dipolar coupling between these neighboring spins and by structure—dependent parameters. A set of differential equations for "average" product operators of the density operator can then be solved numerically to yield the effective size N vs. preparation time. Clear distinctions between the dynam- ics of two- and three-dimensional spin systems are both predicted and observed experimentally.36 A number of three-dimensional solids have been recently ob- served to exhibit a universal growth behavior (when the time axis is scaled by the strength of the square root of the second moments)36 that can be fit to theoretical predictions of an ”incremental shell model; " the effective size increases as the cube (or third power) of the preparation time. In case of a presumed surface film of protons, the experimental results could be fit to the model with an approximately quadratic dependence of the effective size upon preparation time, as predicted 72 for a two-dimensional spin distribution.36 An infinite one-dimensional distribution of uniformly-spaced spins would be expected to provide the simplest experimen- tal test of the various models. We report here results for a close approximation to such a spin system: the 1H nuclei of hydroxyapatite, Ca5(OH)(PO4)3. Since the multiple-quantum coherences are created by homonuclear dipolar couplings, we need only consider the hydroxyl protons (or fluoride ions) in the structure of hydroxyapatite (or fluorapatite). Figure "-1 shows the basic idealized geometry of the ‘H spin system of hydroxyapatite: infinite linear chains of protons having a uniform spacing of 344 pm, with each chain surrounded by six other chains at a distance of 942 pm.“59 The largest intra-chain dipolar coupling is some 20 times greater than the largest inter-chain dipolar coupling, and should dominate the multiple-quantum dynamics. The weak heteronuclear dipolar couplings to the 3‘P nuclei (<2 kHz) can safely be ignored.“4 In order to use apatites as a model for studying one-dimensional MO dynamics, one must have knowledge about the occurrence of interruptions in the 1-D chain, due to either vacancies or substitutions (collectively referred to here as "defects"). Such defects are commonly present at significant levels in both synthetic as well as naturally-occuring apatites. The ability to obtain such information about defects in apatites from MQ NMR experiments would thus provide a useful method for investigating such systems, in addition to better defining the degree of ideality of one-dimensional spin systems in powders and single crystals. In this study we present experimental results on proton MQ NMR dynamics in a stoichiometric hydroxyapatite sample (HAP-M), a hydroxyapatite sample containing defects (HAP-N), and a series of fluorohydroxyapatite solid solutions, Ca5(OH)(1-x)Fx(PO4)3, with the fluoride ion (replacing a hydroxyl group) forming 73 Figure "-1. Schematic idealized arrangement of linear columns of protons (black circles) in calcium hydroxyapatite (Ca50H(PO4)3). The central column is sur- rounded by six neighboring columns. The distance between columns is 942 pm and the distance between intra-chain protons is 344pm. The position of protons in four of the six neighboring columns is actually approximately 260pm below the black circles in the monoclinic form, which exists only for very stoichiometric samples. In the more commonly-occurring hexagonal form, the protons in three of the six neighboring columns are located about 260pm below the black circles due to statistical disorder. The geometry of the fluorine atoms in fluorapatite (Ca5F(PO4)3) is similar to the arrangement of protons in hydroxyapatite, with a distance between columns of 937pm and an identiml intra-chain distance. 74 a 1° —> 942 pm L - 120° 0 O O b 6,1 C5 63 1 Q. i 344 pm Q 75 a defect in the 1-D chain of spins. We model these results for a hydroxyapatite sample with a slight hydroxyl deficiency in terms of a 1-D cluster model described in the Discussion section. The 1-D cluster model considers randomly-distributed defects in apatites as producing a distribution of 1-D clusters of varying lengths, and uses the MO response of a stoichiometric hydroxyapatite sample as a ”cali- bration". In addition, we report 19F MD NMR data on a single crystal of mineral fluorapatite at several orientations in the magnetic field, thereby scaling the dipolar interactions within the chains by a known amount. This approach, when applied to a sample containing few defects, should eventually allow one to scale the rel- ative contributions of one-dimensional and higher-dimensional MO growth, and thus permit better isolation of the effects of differing dimensionality. 2. Multiple-Quantum NMR Dynamics A. Density Operator Description of Multiple-Quantum NMR Dynamics The description of NMR experiments can be started by first considering simple two level systems. An isolated spin-1/2 particle in an external magnetic field has the two eigenstates I15) and I—é), which represent the two allowed orientations of its angular momentum.‘7'6° The state vector, a superposition of these basis states, is given by a linear combination of these two states ' 1120)) = 1,2111%) +c-l/2(t)l — 1) (II-1) where c1/2(t) and c_1/2(t) are the (in general) time-dependent complex coeffi- cients. The state of the system can be described by the values of c1/2(t) and c_1/2(t), which are solved by using the time-dependent Schroedinger equation 111(1)) = 4111111)). ("-2) When the only Hamiltonian is the Zeeman Hamiltonian, the time-dependent coef- 76 ficients are given by Cl/2(t) = a-erp(ia)ea:p(—iw0t) (ll—3a) c-0211) = hemmexpawot) (II-3b) where a2 + b2 = 1, and a and [3 are real numbers representing phases.47 The ex- pectation value of any observable can be calculated from |¢:(t)). Expectation values of the three components of magnetic moment are represented by Mt) = (¢(t)lthzlw(t)) ("'48) AN) = (1(t)|7thI1/2(t)) ("'45) Mt) = (wlt)|7hlzlw(t)) (”'40) which are solved by using the raising and lowering operators defined by 71 = fl(1+1)—m(m:l:1)|m:l:1). (ll-5) The result is 1,“) = 7habcos (a — a + wot) (II-6a) 1,,(1) = 7habsin(a — 13 + 001) (ll-6b) Mt) = 7h(02 — (12)/2. (ll-6c) The expectation values of the magnetic moments of transverse components for an individual spin precess about the external magnetic field with angular frequency (.00 and phase a — 5. However, the expectation value of the magnetic moment of the 2 component is constant, and, (at equilibrium) is proportional to a population differ- ence of the two spin levels, which in turn is proportional to the energy difference in the “high temperature approximation” (Boltzmann distribution). Since the behavior of a single spin or group of spins cannot in general describe a macroscopic sys- tem, we introduce an ensemble of independent two-level subsystems. Figure "-2 shows the mixed and pure states of this ensemble. The transverse components of the subsystems have the same frequency, but need not have the same 77 Figure "-2. Random phases (a) and correlated phases (b) in an ensemble of two level systems. Taken from Ref. 61. 78 phase for a the mixed state (Fig. ll-2a). The average transverse components of the subsystems over all values of the phase is equal to zero, and only the lon- gitudinal component remains (see equation "-6). In a pure state (Fig. Il-2b), the behavior of the macroscopic system is identical to that of each microscopic sub- system, due to each subsystem having the same frequency and phase (phase coherence)!“61 The term “coherence” is defined as the presence of some de- gree of phase coherence between the basis states of the isolated subsystems throughout the ensemble.61 An ensemble of spins can be described by the density operator, which is represented by p = Z pilzbt) (11%| = 111111. ("-7) where p,- is the probability of each state 2/2, occurring in the superposition.62 The equilibrium density operator can be represented by the M) (M,| basis whose matrix formulation of Eq. "-2 is given by p = iii/2C”? il/Qcil/Q . ("-8) Cmaul/2 6—1/20-1/2 With the probabilities from the Boltzmann factor p(M = 1.1,) = e$p(-tho/kT)/Z, (ll-9) the equilibrium density operator is given by PM = exp(—hw0Iz/kT)/Z ("'10) after averaging over the phase differences. From the high temperature approx- imation, the partition function Z is equal to 2I + 1. For the equilibrium density operator of a spin 1/2 system, Eq. "-3 can be rewritten as [p..]= 6$P(-hwo/2kT)/2 0 '1 0 exp(-fwo/2kT)/2 ' The random phase among spins of different subsystems in the ensemble averages (ll-11) the off-diagonal elements to zero, but an application of an appropriate rf pulse to 79 the ensemble results in non-zero off-diagonal elements in the matrix. The exis- tence of the non-zero off-diagonal elements from a single pulse is described as single-quantum coherence. The order n of a coherence between states Ir) and Is) is defined as n = AM = IM. - M,|, (II-12) where M, represent the total (summed) M, values of all the spins in the state |r). The dimension of the density matrix increases as size of the coupled spin system increases. The presence of phase coherences in large spin systems produced by a proper pulse sequence results in non-zero off-diagonal elements that represent the multiple quantum coherences between states. B. Time Development of the Density Operator in the Rotating Frame The density operator determines the‘ state of a system at any time. The de- . velopment of the density operator p is governed by the Liouville-von Neumann equation. 31,42 = 2110.11]. (ll-13) When the Hamiltonian H is time-independent, the formal solution of Eq. "-13 is represented by 90‘) = exp{-(i/B)Htlp(0)eXP[(i/fi)71t]- (ll-14) When p and ’H commute, no evolution of the density operator occurs 90) = p(0)expl-(i/hWtJeXPIG/fimt] = p(0)- (ll-15) If ’H depends explicitly on time, we may satisfy both the Schroedinger equation and the Liouville-von Neumann equation by replacing the propagator that makes the evolution of one state to other state by a unitary transformation in Eq. "-13 with U(t) = Tezp {—1 j 11(1) 41’] (ll-16) where T is the Dyson “time-ordering operator”. Coherent averaging theory63v64 is 80 used to solve Eq. 15. p(t) is always related to p(0) by a unitary transformation in Liouville space so that the length of the vector representing the density operator remains constant as the system evolves. Since the expectation value of any op- erator can be calculated from p,47 = 6°. 88 different MO orders by multiples of the resonance offset frequency Aw. The spectral width of the MO spectrum is the inverse of the t1 increment. The number of orders detected, nmaz, depends on the value of the phase increment Ag) according to the relationship A45 = 27r/nm33. Figure "-5 shows the time-domain MQ signal of hexamethylbenzene generated by the TPPI method. Figure "-6 shows the frequency-domain MQ spectrum of hexamethylbezene obtained using the TPPI MO experiment. During the evolution time, the increase of t1 duration results in a decay of the signal, which causes the frequency-domain peaks of the MO spectrum to broaden for higher orders. The phase-incremented MQ experiment proceeds just as described above, but since the phases of the preparation pulse are incremented by A4), with a fixed evolution period, the t1 domain MO signal does not decay and shows periodicity (shown Fig. "-7). Fourier transformation of the time-domain MQ signal with respect to 42 generates the series of 6-function peaks corresponding to the MO order n. Figure "-8 shows the frequency-domain MQ signal of hexamethylbezene obtained with the phase-incremented MQ experiment. Most MO NMR dynamics in solids are interpreted in terms of the effective sizes that are measured from the integrated intensity of MO orders. The spectrum from the phase-incremented method has a higher signal-to-noise ratio than that from the TPPI method, which reduces the number of transients required. The integrated intensity obtained by simply measuring the peak height also makes the data analysis simpler. Since enough digital resolution can be obtained by replicating the time-domain MQ signal from the phase-incremented method, due to the periodicity of the time-domain of MO signal, the number of time-domain points required in the phase-incremented method is much less than that of the 89 a) 0 7f 27.. b) I 1 I l I . i I) i w I l ,' / 1 i . J 101v 111r 12x 13x 14x 15x 161r —.-—«_.__ -. —-—- __ 0' 1r 21r Sr 41 511‘ Sr 71r 8x91 PM” Figure "-7. Time-domain MO interferogram of hexamethylbenzene using the phase-incremented method and the pulse sequence shown in Fig. "-6. The 60 complex points (21r) in (a) were experimentally obtained, and were replicated up to 480 complex points (160). 90 1le )1, 1.1 JUL? N U @1ij JUL/ wLJLDLJL. 8 18 20 22 24 26 n(nurrbsrofquanta) Figure "-8. Frequency-domain 1H MO spectrum of hexamethylbenzene using the phase-incremented method and the pulse sequence shown in Fig. "-6. All parameters are the same as in Fig. 8, except the fixed t1 interval was set to 200 ns. 91 TPPI method. However, the 6-function peaks of the phase-incremented method do not yield the spectral information about the different frequencies occurring within each order.32 When only the information from the integrated intensity is needed, the phase-incremented method is preferable due to the lower time required to ecllect the data. The pulse programs for the MO pulse sequences used on the VXR 400 spectrometer were kindly provided by Dr. T. Barbara of Varian Associates. We have modified the MO pulse sequences for our experiments, as shown in Appendix A. D. Statistical Model of MO Coherence Intensities In MO NMR experiments, the intensity of each order in the MO spectrum depends on the number of correlated spins (effective size). The dimension of the density matrix is 2N for an N spin-1/2 systems. The number of spins in a macroscopic sample is about 102°. It is impossible to calculate the density matrix of such a macroscopic sample in solids. However, in the statistical mwel31-52-53'55'“, we assume that the intensity of each order in a MO spectrum is proportional to the number of possible MO transitions, and that the transition probability of each order is the same. The effective size can be measured by counting the number of such MQ transitions. ' In a strong Zeeman field, an N spin-1/2 system has 2N stationary states, which can be classified according to the total magnetic quantum number M,, given by M, = z; m,, (II-34) where m,, is the eigenvalue (m,, =‘t1/2) of the i-th spin in the system. The energy eigenvalues of the Zeeman term are represented by (see Eq. l-3) E, = —7hH0M,. (ll-35) 92 The energy levels can be sorted by the number of spins in the la) or |+ {-5) state, no, and in the [13) or | — 1), 123 state, since we can rewrite Eq. "-34 as M, = (no, — n3)/2. (ll-36) The degeneracy S2 of states having a particular energy level E, is represented by n = NI/(naing!) (II-37) Substituting Eq. "-36 and, using the fact that no, + 71.3 = N, we can rewrite Eq. "-37 as 12 = N!/[(N/2 + M,)!(N/2 — M,)I]. (II-38) The energy levels of the Zeeman term and the total magnetic quantum number M, are shown in Figure "-9. Degenerate in the Zeeman energy, the levels within a manifold are shifted and split by the internal interactions, making possible a large number of spectroscopic transitions or coherences. The dashed lines indicate the MO transitions, in which several spins flip together subject to the general rule, 72 = |AM|. The number of all possible transitions as a function of their order, n, can be calculated by combinatorial arguments. The number of n-quantum transi- tions is represented by 2()t) = (1:2) ‘ «139) where (3) = a!/[(a — b)!b!]. The number of zero quantum transitions between pairs of states in the same Zeeman manifold is given by -N/2+l 2° = 2 (11,111) [(m’iM) ‘ 1] = [(211v ) " 2N1 ("'39) M1=N /2-1 For nonzero orders, Eq. "-39 is well approximated by the Gaussian distribution using Stilring's approximation I(n,N) = 4N/\/—N-7r_*exp(—%:) (N 2 6). (II-40) E. Simplified Models of Multiple-Quantum Dynamics The time-dependent behavior of multiple quantum coherences can be -N/ 2 -N/2+1 — . - won/2] 2 EZ o +YhH0N/2 1 +Yf1HO[N/2-1] N +Yf‘lHo[N/2-2] N(N-1)/2 2(N+1I2) +‘Yf1Ho[1/2] (n N)"2 (NH/2) (TC N)1/2 -, Yleo [N/2-2] N(N-1)/2 - YhHo[N/2-1] N - YhHo W2 1 Figure "-9. Schematic energy level diagram for an N (odd) spin-1/2 system in a Zeeman field, and the degeneracy number for $2 of each state. Taken from Ref 53. 94 Increasing Preparation Time (m) —> -> '74; {7.9 ,/ "‘\' "Vr Figure "-10. Symbolic representation of the spreading of multiple spin correlations in a coupling network with increasing preparation time. The circle denote spins and the lines indicate the link of MO coherences through dipolar couplings 95 calculated explicitly using the Liouville-von Neuman equation applied to the den- sity matrix. Since the multiple-quantum coherences in solids are created through homonuclear dipolar coupling, we can rewrite Eq. "-14 as p(t) = eXPl-(i/fimutlm0>6Xp[(i/E)Hnt] (ll-41) where Hg is defined in Eq.lI-31. For short times, an explicit form of the solution of Eq. "-41 can be written as the power series W) = p(0) + (i/h)t[p(0),’Hn] + (i/h)2(t2/2!)[lp(0),7101.710] +(i/h)3(t3/3!)[111401710], 71011710] + '°°- ("'42) The contribution of high order terms in Eq. "-42 becomes more and more impor- tant with increasing time. Figure "-10 shows symbolically the development of multiple-spin correlations by way of the homonuclear dipolar coupling. The num- ber of correlated spins increases with increasing preparation time. The explicit calculation of the MO dynamics using the Liouville-von Neuman equation is limited to small spin systems (3 9 spins) because it is computationally demanding. Reduction of the size of the operator space can increase the size of the spin systems that can be treated. Since classes of coherences are detected in an M0 experiment, as opposed to individual coherences, a simplified, albeit inexact, calculation of MO dynamics is possible by use of an average operator. The following sections will discuss two simplified calculations of the MO dynamics: a directed walk through Liouville space (hopping model)57 and a incremental shell model3‘3 making using of the concept of an average operator. a. Hopping Model The density operator can be represented by a vector in Liouville space, w» = 1:50 fiKggK..(t)lKnp) (II-43) where g Knp(t) is the component of the Liouville states, K is the number of single- 96 spin operators involved in forming the product operator |Knp) of the Liouville state, 72 is the order of the coherence, p labels the different states having the same val- ues of K and n. The symbol for Liouville states, In), is used to distinguish these from the symbol Hilbert states, |---). The equation of motion in Liouville space is represented by a vector equation 1w» = 481711)) (II-44) where ”H is the superoperator [H, ....]. Eq. "-43 can be expressed in terms of the components 9K“, 21179113110“) : —i g: 2 Z QKnP;K’n’p’9K'n’P’(t) ("'45) I n! pl where Kn me) = H411“: 2;, (II-46) and Knp 'H K'n'p’ QIfitp;1("n’p’ = (KanKnp) ("'47) The different oscillatory behaviors of the various components of the density op- erator in Eq. "-44 are hidden by destructive interference and resultant decay, which decrease the number of degrees of freedom. Since Eq. "-45 is a kind of first order kinetic equation, Munowitz etc. assumed that the motion of the density operator can be solved by using the following equation 1 g; = By (II-48) This equation represents a generalized hopping model, in which the elements of R give the rate of change from one component of g(t) to another. R is a matrix of real numbers. For this model, we need to define the space over which the coher- ences hop and to develop the rates and selection rules that govern the motion. The size of the Liouville space can be reduced by grouping together modes 97 n-quantum coherence Figure "-11. Projection of Liouville space onto a two-dimensional plane. Each point corresponds to a family of K-spin/n-quantum operators. Taken from Ref. 57 98 n-quantum coherence .p Figure "-12. Pathways of the growth of MO coherences in Liouville space for a 6-spin system under the 1-spin/2-quantum Hamiltonian. Taken from Ref. 57. 99 of the quantum numbers K and n. In general, the components of the vector 9 are labeled as 9.3,. The projection of the Liouville space is shown in Fig. "-11. Each point on a two-dimensional grid corresponds to a family of coherences ( |Kn) operator basis). K runs from 1 through N and Inl runs from 0 through K. With the assumption that the coherences are of equal magnitude, the number of cperators ggndepends on K, n,and N. The selection rules and hopping rates from one site to the other site through Liouvlle space must derive from the Hamiltonian. The MO pulse sequences in Fig. "-6 used in our experiments create the 1-spin/2—quantum dipolar Hamiltonian which adds one (or subtracts) spins, and changes the order by 2 quanta at a time to a multiple-spin mode. Under this Hamiltonian, the selection rules in the Liouville space are AK = i1, An = 1:2. (ll-49) The detailed proof is in Appendix B. The pathways through the Liouville space may be constructed on the basis of the selection rules in terms of the specified starting point, that the reduced density operator p(0) at thermal equilibrium is proportional to I, (K=1, n=0). For example, Fig. "-12 shows the allowed changes of the Liouville states of a six-spin system under a 1-spin/2-quantum operator. The time development of p an be solved by using the rate equation (Eq. ll- 48). With the assumption that all coherences have equal magnitudes, the hopping rates depend on the degeneracies of the coupled states and the strength of the dipolar interaction. Each element of the rate matrix is given by RKn;K'n’ = WKn;It"n’SI (”'50) where WK”; K1,.) is the generic hopping rate between 9m. and gran). and S, is the lattice parameter, which reflects the strength of the dipolar couplings. The general 100 behavior represented by the statistical factors (the degeneracies of the Liouville state) as well as the individual details represented by the structural factor (dipolar couplings) are of interest in this model. As a first approximation, we will con- centrate on the universal trends, leaving 5', as an adjustable parameter to fit the behavior of arbitrary systems of N spins to specific experimental examples. The lattice parameter, which depends on the structure of the material, may be given by a lattice sum of the coupling constants 51 = 7%, 231193). (ll-51) t= x/nmax. A 2 ms delay after the mixing time is used to allow the transverse magnetization to decay. The data are detected with a 7r/2 pulse, followed by a 100 ps spin locking pulse. A pseudo-10 spectrum with 64 complex points in the "tl-domain" (phase-incremented) was obtained by sampling a single complex point in the t2-domain 35 us after the last (spin-lock) pulse. Unwanted odd order multiple-quantum coherences are eliminated by a 180° phase shift of the detection pulse at every scan. Because of the periodic nature of the data, increased digital resolution was obtained by replicating the "FID". The number of correlated spins N (the effective size) was obtained by least-square fitting of the orders of coherence (excluding 0 quantum coherence) to a Gaussian function, according to the widely-employed statistical modelF“""5453-55-56 The MO intensity distribution should be symmetric, I.1 = I.n theoretically from Eq. "-40. Since the intensities of the positive MQ orders and the negative MQ orders are slightly different experimentally, the sum of the positive and negative MQ orders (In +l.n) was used in the fitting. The sum of all multiple-quantum intensitie [2(lo + In + l...)] at a given preparation time was normalized to unity. Since the chemical shift anisotropy of single crystal fluorapatite (hereafter referred to as FAP) is known (84ppm),‘3‘5 the angles between the c axis of single crystal FAP ( the a” direction) and the external magnetic field were calculated from the observed chemical shifts. The PC-based computer program ANTIOPE66 was used for explicit calculations of MO dynamics of single crystal fluorapatite (FAP) 113 at different orientations. A basic cycle time of 30 us was used. Calculation of spin dynamics of up to 5 spin-1/2 nuclei is possible with the ANTIOPE program. B. Sample Preparation and Characterization Hexamethylbenzene, 06(CH3)6, was obtained from Aldrich and used for the multiple-quantum experiments. The sample of partially monoclinic hydroxyapatite (hereafter referred to as HAP-M) was prepared,67 analyzed and provided by Dr. Bruce Fowler of the National Institute of Standards and the National Institute of Dental Research. Using the intensity of weak X-ray powder diffraction peaks char- acteristic of monoclinic hydroxyapatite59 relative to that of a strong peak arising from the hexagonal form”, along with theoretical calculated intensities for the _ powdered monoclinic form, his analysis of HAP-M showed that roughly 70 % of the sample is in the monoclinic form. Its hydroxyl content is therefore assumed to be highly stoichiometric compared to most hydroxyapatite preparations, espe- cially precipitated samples. Another sample of hydroxyapatite (hereafter referred to as HAP-N) was prepared by aqueous precipitation and characterized by many methods;58'7° a hydroxyl content in HAP-N was determined to be 92% by quanti- tative 1H MAS-NMR, and 81% by IR.68 The solid solutions of fluorohydroxyapatite (FOHAP) , Cas(OH)1.xe(PO4)3, with the different fluorine mole fraction (x = 0.24 and 0.41) were synthesized by aqueous precipitation at a boiling temperature.70 Their characterization has been previously reported.”71 A Specimen of single crystal fluorapatite was kindly loaned by Dr. Bruce Fowler. The color of this sam- ple is pale yellow, with no obvious inclusions, and the diameter and length are about 4mm and 7 mm respectively. 114 4. RESULTS A. 1H Multiple-Quantum NMR Study of Off-Resonance Effects Hexamethylbenzene was used to check off-resonance effects in the MO NMR experiments. The half-height linewidth of hexamethylbenzene in the static 1H spectrum is about 11 kHz. The planar benzene rings of hexamethylbenzene in the triclinic unit cell form a nearly perfect hexagonal net.72 Since fast reorientation of each methyl group along its 0;, axis makes the three proton nuclei equivalent, the CSA of hexamethylbenzene is negligible. The sixfold hopping of the hexam- ethylbenzene about the Cs axis of the benzene ring lessens the intramolecular dipolar coupling between ortho-, meta-, and para-methyl groups.73 Thus, breadth of the peak arises mostly from the intermolecular dipolar coupling. Figure "-18 shows multiple-quantum spectra of hexamethylbenzene with a preparation time of 504 [18 obtained with different transmitter carrier frequencies. The location of the transmitter was varied from the middle of the 1H static lineshape to 4 kHz off-resonance in intervals of 1kHz. The 2 quantum peak on resonance (Fig. Il-18a) is more intense than any other peak, but moving the transmitter from the center of the spectrum diminishes the relative intensity of the 2 quantum peak, eventually resulting in an inverted peak (Fig. ll-18b to e). The intensities of the higher-order peaks, shown in Fig. "-18 with the same vertical scale, also decrease with increasing resonance offset. Static 1H spectra, normalized to the same peak height, of HAP-M, HAP-N, FOHAP x=0.24 and FOHAP x=0.41 are shown in Figure "-19. The spectrum of the low specific surface area (2.1 m2/g) HAP-M sample in Figure lI-19a does not show evidence of a peak from surface adsorbed water (around 5.6 ppm)“‘9 but instead only a 2 kHz broad peak, whose width is due to the CSA of the hydroxyl C) b) ...JLJLJ a) JUL.) 2 4 6 8 10 12 14 16 n (chquanta) Figure "-18. 1H PI-MQ—NMR spectra of hexamethylbenzene with different trans- mitter carrier frequencies. See text. 116 ., J /\\ J/ “L IYVYVITYY'IYTTTITTTTTTIVWYYITIITTrIIITI[TITTrIT1llTllllllTTifilV[III'IT—Y17] 7O 60 50 40 30 20 10 O -20 -40 -60 00111 Figure "-19. 1H 1-dimensional spectra of apatite samples. HAP-M (a), HAP-N (b), FOHAP x=0.24 (c), and FOHAP x=0.41 (d). 117 group, and homo— and heteronuclear dipolar couplings. The strong signal in HAP- N around 5.6 ppm (Fig. ll-19b) has been assigned to mobile water at the surface. The peak of hydroxyl groups in HAP-N is concealed by that of surface adsorbed water (the specific surface area of HAP-N is 37m"’/g).68 The weak intensity of the 1H MAS-NMR spinning sidebands of the surface absorbed water groups indicates that homonuclear dipolar couplings among these protons are negligible.69 Figure "-20 shows the 1H multiple-quantum spectrum of the highly stoichio- metric HAP-M sample for a preparation time of 864 us, as well as the Gaussian fit corresponding to an effective size of 12.3. The higher orders of multiple-quantum coherences deviate from a Gaussian distribution,31 being more intense than pre- dicted. Figure "-21 displays the intensities of the multiple-quantum peaks for HAP- M with a preparation time of 864 ps as a function of resonance offset. The isotropic chemical shift of hydroxyapatite is 0.2 ppm, but the position of highest intensity of HAP-M (2.8 ppm) is regarded as “on-resonance”, and the transmitter offset is varied by up to :2 ppm (800 Hz at 9.4T). The change of the transmitter offset position has only a slight influence on the intensities of the 2 quantum and 4 quantum peaks, shown in Fig. ll-21a with the same vertical scale. Since the effective size is measured from the ratios of intensities of multiple-quantum peaks, we can redraw Figure Il-21a with intensities normalized to the 2-quantum peak. The resulting intensity profiles, shown in Figure II-21b are not changed much by varying the transmitter offset by 12 ppm for a preparation time of 864 [18. Figures ll-22a and b have the same parameters as Fig. 21 a and b except for the preparation time (1872 [18). The intensities (with the same vertical scale) of multiple-quantum coherences obtained on resonance are always stronger than 118 \ \ \ . ,i .72; .\ s l \ é " . \\ Ii \ c: \ ‘ " \ n (number of quanta) Figure "-20. 1H MD NMR spectrum of HAP-M using a preparation time of 864 us. The dashed line represents a Gaussian fit to the data with effective size N = 12.3. 119 1.2 lb) 01.0‘ ' N I 908‘ E goe' 20.4: . '7: 502‘ L5 1 I 0.0 2,-ssr-!2;__i 93300 S a); a. l a 3 52001 3 g 4 £1001 a I '5 3 n o e.f.rtr!vl-———I o 2 4 6 81012 MQorder figure “-21. Intensities of the ‘H PI NMR MO orders of HAP-M for various resonance offsets using a preparation time 01864 #8. a) Intensities on the same vertical scale. b) Intensities normalized to the 2-0 intensity. On resonance (open square), 1 ppm off resonance ( open triangle), 2 ppm off resonance (Open circle), -1 ppm off resonance ( black triangle), and -2 ppm off resonance (black circle). 120 1.2 . b) O 1.0 ‘ ' A N i O O "' 0.8 1 a g I f“ 0.61 E . i 2 041 > . a ‘ a 5, 0.2 ‘ a 5 . 0.0 - + SE 300 S a) 2‘ g 3 :5 200‘ o 3 9 > O '5 ‘ . . 5 ,§ 1001 ‘ ‘ a O 5 , . e 3 ‘ I o - . - . e . - 2 - . - O 2 4 6 8 1O 12 MQorder figure "-22. Intensities of the ‘H Pl NMR MQ order of HAP-M for various resonance offsets using a preparation time of 1872 us. .a) Intensities with the same vertical scale. b) Intensities normalized to 2-Q intensity. On resonance (open square), 1 ppm off resonance (open triangle), 2 ppm off resonance (open circle), -1 ppm off resonance ( black triangle), and -2 ppm off resonance (black circle). 121 those obtained off resonance to the upfield side, but are stronger for the 2 and 4 quantum coherences, weaker for the 8 and 10 quantum coherences than those obtained off resonance to the downfield side. It is difficult to obtain an effective size fitted by using a Gaussian function from the intensity profile for an off-resonance transmitter offset at longer preparation times. Thus, the effective sizes obtained using Gaussian function with different resonance offsets are significantly different from those obtained on-resonance, due to the distortions of the MO peaks. Such resonance offset effects complicate the interpretation of MO dynamics. B. 1H Multiple-Quantum NMR of Hydroxyapatite and Fluorohydroxyapatite Samples The normalized intensities of the various orders of MO coherence as a function of preparation time 1' for the various apatite samples are plotted in Figures ll-23a, b, c, and d. Figure Il-23a shows that the normalized intensity of the zero-quantum coherence of HAP-M decreases, and that of the 2-quantum coherence increases first and then decreases with increasing preparation time; the higher quantum orders steadily grow within this experimental range of preparation times. The normalized intensity of the zero-quantum coherence of HAP-N in Fig. ll-23b shows a different behavior compared to that of HAP-M; it decreases slightly at earlier preparation times and levels off at longer preparation times. Although the normalized intensities of the high quantum orders of the HAP-N are less than those of HAP-M, they present the same behavior with respect to increasing preparation time as HAP-M. The normalized intensities of the zero-quantum coherence of FOHAP x=0.24 and x=0.41 shown in Figure ll-23c and d show a slight variation, and are higher than those of HAP-M and HAP-N. 122 1.0 l o o 0.8 o 0 ° ° 0 o o o b) 0.6 1 0.4 1 g . 2 0.2 g ‘ ‘ ‘ A ‘ ‘ t l ‘ ‘ ‘ " 0.0 .__._-_,_a_n._fl_i_L_l_L E f, 1.0 S a) 2 0.8 1 . , O O '0 0.6 ‘ c . ° ° 0 o o 0.4 ‘ 1 A ‘ ‘ ‘ . ‘ ‘ 5 a 0.2 < ‘ a a D O D O a . I 0.0 ,fl Ll—I—l—l—L.'—._ O 500 1000 1500 Preparationtimems) figure "-23. The normalized intensities of even-order MQ coherences of apatite samples as a function of preparation time. a) HAP-M; b) HAP-N; c) FOHAP x=0.24; d) FOHAP x=0.41. 00 (open circle), 20 (open triangle), 40 (open square) 60 (open circle), and 80 (black square). Nomalized intensity 1.0 0.8 0.6 0.4 ‘ 0.2 0.0 1.0 0.8 0.6 0.4 0.2 0.0 123 ‘1! I I LI 500 1000 1500 Preparation time (us) 124 The effective size obtained from fitting MQ intensities to a Gaussian is useful for studying MQ dynamics. Figure "-26 shows the experimental effective sizes N vs. preparation time for HAP-M and theoretically calculated curves for 1-D and 2-D growth using the incremental shell model. The experimental plot is linear in its early preparation time development, only deviating upward at longer prepara- tion time (where one expects the weaker inter-chain dipolar couplings to become more influential). The solid line represents the calculated curve for a one dimen- sional chains of 1H spins in hydroxyapatite. The value of nn in the incremental shell model was chosen to be 1, since there is only a single uncorrelated spin adjacent to each spin at the coherence boundary. The value of no, (related to a system's dimensionality) was set to 2, representing the total number of adjacent spins outside of the coherence (the two “end Spins” in a linear chain). The dipolar coupling D12 between a pair of nuclei is rewritten from Eq l-50 in the Part I as 012 = (72h) (1 — 3cos2 012)/2ri32 (in HZ). (ll-77) For two neighboring infra-chain protons 344 pm apart (the distance in hydroxyapatite”) and parallel to the external field, D12 = 2.957 kHz, which is 1/3 of the Pake doublet splitting that would be obtained in a 1H spectrum. We used in the simulation an average 012, 1.32 kHz, whose value is the root mean square value of D12 over all powder orientations. Since the value of P2(c03912), (1 — 3c082 912), is averaged out over all orientations but that of P2(c030)2 is not, an average D12 is obtained from the square root of an average sz. This follows the approach used by Levy and Gleason.36 For a 1-D system, a more reasonable approach is to sum the powder-weighted responses foe each orientation of the 1-D chain. The experimental slope is three times larger than the predicted slope. The dashed line shown in Fig. “-24 is the calculated curve for a two dimensional 125 50 I 40 1 ‘ I 2 .. I g 30 m 4 I .3 ‘87 20 . ' n‘i ‘ I I 10 ' I I ..«v 4 I ...v" . ~G""‘ o ‘ "'1'"-' fl . 0 500 1000 1500 2000 Preparation time (us) Figure "-24. Effective size N versus MQ preparation time for HAP-M. The solid line is a theoretical calculation using the incremental shell model for a one-dimensional spin system with the appropriate dipolar coupling strength and bo adjustable parameters (see text). The theoretical calculation of a two-dimensional system is represented as the dashed line. 126 Fe 4 l' 3 cl L- g 1 S 2 d 1 f t W t ' 1 2 3 4 In (mg) Figure "-25. In(N) vs. In(r/Tc) for all of the samples of hydroxyapatite and fluorohydroxyapatite. Solid lines indicate the slopes of data points that are obtained using least squares fitting. HAP-M (open circle), HAP-N (open triangle), FOHAP x=0.24 (open square), and FOHAP x=0.41 (black circle). 127 system assumed to consist of the Six nearest protons in adjoining columns in the hydroxyapatite structure. The calculation was performed by using an inter-chain nearest-neighbor powder-average dipolar coupling D12 of 64 Hz. The value of nn was set to 6, the number of planar neighbors, and n, was chosen to be 4M"? The effective size N grows as some power of the preparation time 7' (N 0( 7"). It is convenient to plot In(N) versus In(r/rc) to obtain the growth exponent a as shown in Figure "-25 , where To is the basic cycle times3 The growth exponent for effective size vs. preparation time is equal to the slope of this plot. The MO data for HAP-M and HAP-N show bi-exponential characteristics. The slopes of HAP-M and HAP-N for short preparation times (shorter than 864 us) are 0.98 (correlation coefficient = 0.988) and 0.83 (correlation coefficient = 0.998), and those for long preparation times (longer than 1008 ps)are 1.78 and 1.82 respectively. FOHAP =0.24 and FOHAP x=0.41 have single exponential characters. The slopes for FOHAP x=0.24 and FOHAP x=0.41 are 0.54 (correlation coefficient = 0.963) and 0.50 (correlation coefficient = 0.936) respectively. The decays of the absolute intensity of the sum of all multiple-quantum coher- ences of the apatite samples are shown in Figure lI-26a as a function of prepa- ration times. The decay function corresponding to irreversible relaxation can be assumed to be exponential: |(1') oc 6Xp(-1'/Td). (ll-78) The different relaxation times Td of the different MQ coherences can be distin- guished by taking the logarithm of Eq. "-78. Figure ll-26b shdws a plot of ln(l(r)) vs. the preparation time 1'. Since the Td relaxation time is the inverse of the slope in Figure II-26b, we have offset the intercept in order to see the slope more clearly. The Td relaxation times of HAP-M, HAP-N, FOHAP x=0.24, and FOHAP 128 In(intensity) A //‘/H fl"? 2 - r e . 2500 l a 2000 . ‘ 4 I g 1500 1 2 3 ‘ ' o 1000 1 ‘ o I 1 ‘) I: g o o 500 . ‘ . : c o l ‘ a 2 o f r f 0 500 1000 1500 Preparationtlmeots) figure "-26. Decay of absolute MQ coherence intensity of apatites. HAP-M (open circle), HAP-N (open triangle), FOHAP x=0.24 (open square), and FOHAP x=0.41 (black circle). a) Intensity vs. Preparation time. b) In(intensity) vs. preparation time, arbitrarily offset in the vertical direction. 129 1.0 a) A A 1 A 00 .- 2 . T 0.8 “ I : . . g g! A O 1 a . ‘ ' ‘ .a . 1 8 :C: 0.7 v ! ' I ' f r - .. E 0.3 l:- as o < o E b) A A 0 0.2 ‘ o a. o 8 O O o O A 20 . A 0.1 A A A A A 0.0 ~ . a r f v . . . 0 200 400 600 800 1000 Preparation time (us) figure "-27. The normalized intensities of 00 and 20 coherences of a single crystal mineral sample of FAP vs. preparation time at twoorientations with respect to the external magnetic field. a) 00 at 90° (black circle) and 00 at 62° (black triangle). b) 20 at 90° (cpen circle) and 20 at 62° (open triangle). 130 1' ! i 9.0 , g ...r .5 ¢ . § 80 § 0 G g . 7.0 - . - . - . . , - 0 200 400 600 800 1000 Preparation time (us) Figure "-28. In(absolute intensity of MO orders) vs. preparation time of single- crystal fluorapatite at two different orientations. 90° orientation (open circle) and 62° orientation (open triangle). 131 x=0.41 are 742, 601, 615, and 610 ps respectively. C. 19F Multiple-Quantum NMR of a Single Crystal of Fluorapatite Figures II-27a and b show the intensity of the zero-quantum and two-quantum coherences of a single crystal FAP at two orientations ( c axis making the angles 62° and 90° with respect to the external magnetic field). Unlike a powder sample (see fig. "23), an oscillatory behavior is observed for the zero and two quantum intensities. The frequency of the oscillations of the MO coherences for single- crystal FAP is proportional to the dipolar coupling. The dipolar coupling infra-chain of single-crystal FAP at 90° is approximately 2.95 times larger than that at 62°. The logarithm of the absolute total multiple-quantum intensity for the single- crystal FAP sample at two orientations vs. preparation time 7' is shown in Figure "-28. The Td relaxation times of this single-crystal mineral FAP are 694 ps and 645 #8 at 62° and 90° respectively. 5. DISCUSSION A. Effect of Resonance Offsets in the PI-MO NMR Pulse Sequence on the Formation of Multiple-Quantum Coherence Figures "-18, 21, and 22 show the effect of a resonance offset on multiple- quantum coherences. We can give a possible explanation for this effect. In the rotating frame, when only the static magnetic field is present, the effective field is represented by the vector equation in = 1% — an (II-79) where Ho is the vector of the main static magnetic field and the magnitude of 63 is the angular frequency of rotation of the unit vector at rf carrier frequency. We ‘F 132 figure "-29. The effective magnetic field if,” in the rotating frame in the presence of an applied rf field 171. The effective magnetic field is the vector sum of the applied static field 170, the fictitious field ..3/7. and the rf field r‘il. 133 can express the effective field in the rotating frame in the presence of an rf pulse as the vector sum of fig — iii/7 and H1 (see Figure "29) 1;,” = (H‘o an) +1111 (II-80) where E is the radiofrequency field. The magnitude of the effective field is rep- resented by Heff = [(Ho —w/1)2 + H1211” l/2 = (1/1) [(7110 - w)2 + (71102] (II-81) = (1/7) [wa - w)? +wi] ”2 where wo — w is a resonance offset frequency and w = —71H1 is the radiofre- quency field strength. The Hamiltonian of the radiofrequency along the x direction in a frame rotating at the carrier frequency can be represented by 11,, = —71H1I,. (ll-82) A 90° pulse along the x direction flips the initial magnetization (M, = 7Iz) into the y direction (M, = 7Iy) on-resonance. However, the magnetization of off- resonance peaks after a 90° pulse is not in the y direction. As mentioned in the background section, the average dipolar Hamiltonian generated by the 8-pulse cycle with the proper delays (Pl-MO pulse sequence shown in Fig. "-4) is a driv- ing force for creating the multiple-quantum coherences. The effectiVeness of the average Hamiltonian is determined by the preciseness of the 90° pulse length. An increase of the resonance offset results in an increase of the difference between 1?,” and if; (an imperfect 90° pulse when an off-resonance). The system evolves under the influence of the internal Hamiltonian alone after the 90° pulse. The time development of the system in the absence of relaxation is represented by 100) = exp(-iH.ntt)Iyexp(iH.-mt)- (II-83) 134 The average double-quantum Hamiltonian Hyx Is used instead of the internal Hamilonian Hm in MO NMR experiments. There is no resonance-offset cor- rection to zero order for the average double-quantum Hamiltonian used in our experiments,40 but higher order terms quickly contribute to the resonance-offset effects. Even if the pulses are “hard” 6-pulses that perfectly rotate I, to I, for all offsets, the existence of the off-resonance term AwI, changes the time develop- ment of the density operator in Eq. "-83, due to the resonace-offset correction in some order of the average double-quantum Hamiltonian as time increases. The extent to which resonance-offset effects influence the MO spectrum is sample dependent. The 1H CSA of hexamethylbenzene is nearly zero due to the fast reorientation of the methyl groups, and that of hydroxyapatite is estimated to be roughly 14 ppm.74 The placement of the transmitter carrier frequency away from the isotropic chemical shift for hexamethylbenzene shows a severe destmctive interference of the multiple-quantum coherences for a preparation time 01504 ps. The 2-quantum order seems to be more influenced by resonance offset effects than are the higher orders. As resonance offsets increase, the intensities of the 2- Q coherence become weaker and more out-of-phase. To illustrate the magnitude of off-resonance error in the 1H MO spectrum of powdered HAP-M, the MO spectra of HAP-M was recorded with the transmitter positioned at 2ppm (800Hz) upfield, and in a separate experiment 2 ppm downfield, of the highest position in the pattern. For a preparation time of 862 [18, the MO spectra show no distortion, but for a preparation time of 1872 ps the spectra were distorted due to the cumulative effect of off-resonance terms. B. Dimensionality Effects in the Multiple-Quantum NMR of Hydroxyapatite Different samples show different growth curves of MO coherences with in- 135 creasing preparation time. However, it has been demonstrated by Gleason et. al. that when the preparation time is scaled by the square root of the second moments of the individual samples, samples having the same dimensionality show universal MO growth curves.36 Therefore, the growth of MO coherences is related to the dimensionality.36 The study of dimensionality using MO NMR has been performed with two- and three-dimensional spin systems. Hydroxyapatite is a good model for studying one-dimensional spin systems since it has linear chains of uniformly- spaced proton spins, fairly widely separated from each other. The growth of MO coherences for HAP-M at preparation times up to 864 ps in Fig. "-24 is linear as is the calculated one-dimensional growth in the incremental shell model. In qualitative agreement with the calculation for two dimensional growth, an upward curvature in the experimental curve at longer preparation times may reflect growth to other columns of spins. Lacelle53 has discussed the possible significance of the different growth exponents observed for various spin systems. The growth exponents of 1- and 2-dimensional spin systems obtained from the incremental shell model prediction after the initial induction time are 1.00 and 2.05 respec- tively. The bi-exponential character of the curve for HAP-M shown in Figure "-27 suggests a growth of MO coherences with a change in the dimensionality. The growth exponents of HAP-M (0.98 for preparation times shorter than 864 ps and 1.78 for times longer than 1004 ps), which are close to those predicted by the incremental shell model for 1- and 2-dimensional growth, show a 1-dimensional character at short preparation times arising from intra-chain dipolar couplings, and a higher slope at longer preparation times arising from the more 2-dimensional spin system corresponding to inter-chain dipolar couplings. Of course, the repre- sentation of the inter-chain growth as 2—dimensional is inexact, since it has some 136 three-dimensional character. Further more, when the intensities of the orders of HAP-M are fit to an exponential, the fit is somewhat improved over a Gausian fit, and the parameter characterizing the exponential fit is observed to increase linearly at all preparation time. Thus, the effects of different dimensionalities will require further study, possibly by using powdered fluorapatite samples. C. One-Dimensional Cluster Model Clustered samples give rise to a strong intensity of the zero quantum peak compared to non-clustered samples.75 From figure "-23, the strong normalized zero quantum intensities of HAP-N, FOHAP x=0.24 and FOHAP x=0.41 compared to that of HAP-M qualitatively demonstrate the existence of vacancies and/or fluorine ion substitutions that make clusters in a one-dimensional chain. The growth exponent of an infinite 1-dimensional chain theoretically predicted by . the incremental shell model is 1.00. The growth exponent of HAP-N at early preparation times (0.83) clearly shows evidence for a hydroxyl deficiency. The lesser slopes of FOHAP with increasing mole fraction of fluoride ion in fig. "-25 indicate that the FOHAP samples having higher mole fractions of fluorine contain higher defect densities. Thus, MO NMR shows evidence of the interruption of hydroxyl groups in apatite samples. figure "-30 shows hydroxyl groups in an apatite chain segmented by vacan- cies and/or fluorine ions. The run number of a group of hydroxyls is defined as the number of contiguous hydroxyl groups between two “defects” on either side, which can be either vacancies or fluoride ions. Since MO coherences are created by the homonuclear dipolar interaction, the existence of vacancies and/or fluoride ions between hydroxyl groups in an apatite chain hinders the propagation of MO coherences. In order to compare experimental MO spectra with calculated MQ 137 spectra for given defect densities by the statistical model, we assume the follow- ing: 1. defects are randomly distributed (ideal solid solution); 2. MO coherence growth occurs only along the linear chains; 3. MO coherence do not grow across vacancies or fluorine atom substitutions; 4. MO results for stoichiometric hydroxyapatite (HAP-M) yield effective size for each preparation time 7'. Since the sum of the intensity of all MO orders in the absence of decay of MO coherences is the same regardless of preparation times, it can be normalized to one. The intensities of MO orders, In(r), can be represented using the normalized factor74 from Eq. "-39 as In(r) = CZn/4N (ll-84) where C is equal to 1 for a non-selective experiment and 2 for an even-selective experiment. Since the intensity of each order of MO coherence depends on the run number and the mole fraction of spins present in a given run number, the intensity of each order of coherence for a defect-containing sample can be calculated by summing over the contribution from each run number combined with a Gaussian distribution, as given by the following equation 6 _ mm) = 2 ”(13:23) (4M) 1+ for N, s 6 r=2 27 x,(N,1r)_l/2e$p(—n2/N,) for N, > 7. (II-85) where x, is the mole fraction of spins in a run number, and N, is set equal to the run number for run lengths up to the effective size. The value of N, is set equal to the run number when the run number is less than effective size; otherwise, N, is set equal to the effective size. 138 C . . 2 OH O H . | I x x :F or Defect H _ O I _ F 436pm F . Run number=1 :l 2129'“ _ H _ H x | | . O O . I II . Run number=5 O ' III e — 3 203m l?) x F x F . III figure "-30. Arrangement of hydroxyl groups separated by vacancies and/or fluoride ion substitutions in apatite samples. The distance between hydroxyl groups and fluoride ions has been obtained by NMR studies.75'77 The substitution of fluoride ions gives three configurations (l-Ill) along the crystallographic c axis in FOHAP. The dotted lines denote hydrogen bonds. 139 D. Estimation of Defect Densities Using the 1-D Cluster Model. Even if the MO intensities of experimental data obey a Gaussian distribution perfectly, the theoretical MO intensities of defect-containing samples obtained from Eq. "-85 may not be fitted very well by a Gaussian distribution due to the summing of contributions by different run numbers. Therefore, the comparison between theoretical and experimental 4Q/20 intensities for a given preparation time is more relevant than a comparison of effective sizes. Experimentally, since MQ intensities do not correspond to a Gaussian distribution exactly due to strong intensities of the higher order peaks, we recalculated the effective size of HAP-M using the 40/20 intensity at each preparation time. Since the contribution of MO intensities from a 2-D spin system complicates the calculation of defect density at longer preparation times, we use MO data from the linear portion of the curve corresponding to 1-dimensional growth in Fig. "-24. The mole fraction of the run numbers in a 1dimensional chain can be calculated by percolation theory.78 Figure "—31 shows run number distributions for various defect densities. The center of mass of a given run number distribution moves to a larger run number value as the defect ratio decreases. The result of calculations based on this 1-D cluster model and experimental data is shown in figure "-32. The calculation program is given in Appendix 1C. The HAP-N from MD data are closer to the curve calculated for an 8% defect density (that obtained from 1H MAS-NMR) than 0 that calculated for a 19% defect density (that obtained from IR). The experimental curves for the FOHAP samples are lower than the calculated curves. There are several possible explanations for the deviation between the calculated and experimental curves for FOHAP samples. First, hydroxyl vacancies exist in the one dimensional chains of precipitated a? 140 apatites (compare the difference between HAP-M and HAP-N). These vacancies in the hydroxyl chains can create additional defects in addition to those created by fluoride substitutions, and could explain the discrepancy between the calculated and experimental curves of FOHAP. The experimental curve of the FOHAP x=0.24 sample is close to the theoretical curve for a 41% defect density corresponding to a 22% hydroxyl deficiency. Secondly, if fluoride ions substitute randomly for hydroxyl groups, the molar ratio of the configuration III to configurations l or II shown in Figure "-30 for FOHAP x=0.24 is 0.32. Yesinowski et. al. have shown that the peak arising from the configuration III is not detected with 1H MAS-NMR.69 If configurations l and II are more preferable than configuration III, the decrease of the number of fluorine ions in a row increases the mole fraction of small run number. Therefore, non-random substitutions of fluoride ions in FOHAP can result in the lower experimental intensity compared to the calculated one. Thirdly, the eight-pulse MQ sequence for solids shown in Fig. "-6 ideally eliminates heteronuclear dipolar couplings and J couplings of 1H and 19F.75 In these experiments, we have given applied the MO pulse sequence at the proton frequency. The effectiveness of the decoupling is determined by the basic cycle time, the time scale of mutual spin flips of the irradiated spin system T2” and the strength of the heteronuclear dipolar coupling between 1H and 19F. A heteronuclear dipolar coupling between two different spin 1/2 nuclei is represented by D15 = (7n,h)(1 — 3cos2 912)/3r?2- (ll-86) The 1H-‘9F dipole coupling in configuration I parallel to the external magnetic field is equal to 7.27 kHz. The isotropic chemical shift of hydroxyl groups not adjacent to a fluoride ion is 0.2 ppm and those of configurations l and II are 1.2 ppm and 141 Run number figure "-31. Run number distributions for various defect densities in .a one- dimensicnal chain. The mole fraction of spins in a given run number is calculated by percolation theory"; 8% (open circle), 24% (open square), and 41%( cpen triangle). 142 0.4 0.3 ‘ ° 2 .6 I 5 0.2 « I .5 t 3 0.1 « 0.0 . . . . . , 200 400 600 800 Preparation time (us) figure "-32. 1-D cluster model of MO growth for various defect densities in apatite, compared to experimental data. Lines for the various defect densities are calculated using Eq. "-84. From top to bottom, lines correspond to 8%, 19%, 24%, and 41% defect densities. HAP-M (open circle), HAP-N (open square), FOHAP x=0.24 (open triangle), and FOHAP x=0.41 (crosses). 143 0.4 0.3 r f 2: '75 1 C 2 .5 a CS, 0.2 v 1 ------ "'" 0.1 - ....... ‘5 ....... A a" ........ .— ------- ‘ A .... 1'- ---- ' firzzzzrsisz“‘ I I 0.0 . r - r ' . 200 400 600 800 Preparation time (us) figure "-33. Comparison between experimental 40/20 intensities and calculated 40120 intensities for FOHAP samples after excluding the two hydroxyl groups adjacent to fluoride ion. From top to bottom, lines correspond to 24% and 41%. FOHAP x=0.24 (open triangle) and FOHAP x=0.41 (black square). 144 1.5 ppm respectively.69 However, the proton adjacent to a fluorine atom will be split into a Pake doublet. This splitting produces a resonance for these protons of typically several kHz, sufficient to prevent them from effectively participating in normal MO coherence growth. Thus, we assume that the protons adjacent to a fluorine atoms do not participate in MO coherences. Figure "-33 shows a recalculation of the 4Q/20 intensities for 24 % and 41 % defect densities using Eq. "-85, when terminal OHu-F groups in a run are excluded. The experimental 40/2Q intensities of FOHAP x=0.24 and FOHAP x=0.41 in Fig. "-33 are closer to the calculated 4Q/20 intensities than those in Fig. "-32, suggesting that this resonance offset effect may be at least partially responsible for the initial discrepancy. Finally, a different coherence in the decay time of different run numbers of spins and/or the different MO order might result in a . difference between calculated and experimental 4Q/20 intensities. However, the logarithmic plots in Figure Il-26b do not show two different slopes for a the decay time within experimental limits, which means that different MO coherences do not have different decay times. D. 19F Multiple-Quantum NMR Dynamics of Single Crystal Fluorapatite. The oscillatory behavior of MO dynamics in a liquid crystal has been previously shown experimentally and theoretically for finite spin systems.5‘f56 Munowitz has theoretically predicted such oscillations for small oriented linear arrays of uniformly spaced spins under the Hamiltonian 71,, where Hz: = 1/3 ED141211“ + Iszzlc)- ("'87) The theoretically predicted oscillations are periodic for two spins but are damped for larger arrays (very small oscillations for arrays of 6 spins).5o The oscillatory frequency of the intensities of 1 Q and 20 coherences for a 2 spin system under 145 1.0 ‘17 1’6 . 0e b) °° 0° c o co . e 0.8 J °° o c ‘I‘ O O 0.6 ‘ 0° .‘ ‘9‘ c o ‘A l o 2 2 4 -l ‘ 0 0° ‘ A 0° 0. ‘ .. Cocoa ‘ ‘ f“ g: ‘ A A ‘ g 0.2 1 ‘ A A E 1 ‘ ‘A A. - A ..A - f 2 , - T , ‘Ap‘ - . E 0.0 _ a 1.0 "In; ‘In‘ :6— _ g do a) ‘. ‘A °° L" 2 , o 9 A ° 0.8 o I ‘ O f O I ‘ O 4 0 ‘ O 0.6 ‘ 0. 3° ‘0 9‘ 0.4 1 A . ° . l I O O ‘ A o 0.2 ‘ A a, o ‘. ‘ A. 0 Do ‘A 0.0 «ll-.9- ' v V ovowo v r - f ‘L—l O 200 400 600 800 1000 1200 Preparation time (us) figure "-34. Simulated MQ dynamics of 2, 3, 4, and 5 spin systems under 11,, for an oriented linear chain using ANTIOPE66 (see text). a) 2 spin system; b) 3 Spin system; c) 4 spin system; d) 5 spin system. 00 (open circle) and 20 (open triangle). Normalized intensity 146 1.0 o < 0o d) 0.8 ‘ O . ‘ O 0.6 ‘ o “ ‘5‘ ¢O°°°O°° 0° 1 1 6° °°22° e 0 ‘ 5“ ‘A 0.43 ‘ 0 ° ‘AAA“ A ‘ 4 °oe¢ A 02. ‘ ‘f l “ 0.0 a e . , .. e - 1 0 f 1 on C) ‘A‘O“ A 0.8 ‘ o ‘A 9 A‘I A J ‘ ‘A 0.6 ec‘A . 00.00 ‘I ‘ Ag 0 5‘“. e 0.4 1 9 ° 0 ° . 00000 O . . e 0.2 ‘ oo o f “ 000.0 0.0 ‘FL v T r I V v v I v 1 f O 200 400 600 800 1000 1200 Preparation time (us) 147 71,, is (1/3)D12. Figure "-34 shows the simulated MO dynamics for oriented chains of 2, 3, 4, and 5 spin systems under the Hamiltonian 71,, as a function of preparation times using ANTIOPE.66 The distance between nuclei is 344 pm; the same as that in fluorapatite, and a 62° orientation with respect to the external magnetic field was chosen for simulations. The MO dynamics of the 2 spin system ( Fig. lI-34a) are both oscillatory and periodic. The dipolar interaction between two spins (d12 = 344pm) is 442.6 Hz, and the periodicity of the MO dynamics for two spin system is 1129.6 ps [1/(2'442.6 Hz)]. An increase in the size of the spin system results in a apparent disappearance of periodicity in the MO dynamics. This is likely due to the destructive interference of the many frequencies present when there are a larger number of unequal dipolar couplings present in a spin system: the actual periodicity may occur at time intervals too long to be revealed by the simulation. The MO dynamics of a 5 spin system ( Fig. ll-34d) shows the oscillatory behavior, but no apparent periodicity. The theoretical calculation of MO dynamics under even different Hamiltonians (71,, and H,,) show the periodicity and oscillatory and periodic behavior for both 2 and 3 spin systems. The oscillatory behavior of MO dynamics in single crystal fluorapatite appears to be damped and periodic. We have no independent measurements of the defect density in this particular crystal of mineral fluorapatite, although mineral fluorapatite generally contain defects caused by substitutions such as hydroxyl and chloride ions. We can attempt to model the observed MO dynamics of this defect-containing sample somewhat along the lines of the 1-D cluster model previously described. However, the intensities of the various orders will be obtained in a different manner. For each run number, we will use the 148 intensities of the individual orders of MO coherence calculated from the ANTIOPE simulation. The overall intensities of a given order n at preparation time 7' i given by following expression Jinn) = 22 err(n,r) (II-88) where r refers to the spin system orf-size r. The obvious limitation of this approach is that run lengths < 5, which are expected to yield high-damped oscillation, are ignored, since no calculations are available. Figure "-35 shows the mole fraction of spins in various run numbers from 1 to 5 and the sum of the mole fraction with different defect densities obtained by using percolation theory.78 For defect densities larger than 40%, the sum of the mole fraction of spins in runs 3 5 is about 0.8. Therefore, the contribution of runs having more than 5 spins to the MO dynamics for defect densities larger than 40% is small. However, for smaller defect densities, the larger runs will dominate MQ dynamics. Nevertheless, since these runs will exhibit more damped behavior, we can hope to reproduce the essential features of the experimentally-observed oscillations. Theoretical intensities of MO coherences for an oriented spin system are calculated by using Eq. "-88. Figure lI-36a and b show a comparison between the experimental 20 intensities of single crystal FAP, and calculated 2Q intensities for assumed 30% and 50% defect densities at two different orientations (62° and 90° with respect to the external magnetic field) as a function of preparation time. The calculated 20 intensifies for both 30% and 50% defect densities at the two different orientations show similar oscillatory behaviors. Although the defect density of single crystal mineral fluorapatite cannot be reliably estimated, the agreement does not appear to improve as the defect density decreases from 50% to 30%, especially in the positions of the maxima and minima in Figure Il-36a. A defect 149 density less than 30%, which is quite reasonable, would appear to yield even better agreement if the same trend occurs. 19F MO NMR spectra of single crystal FAP (not shown here) at two orientations (62° and 90°) display 60 peaks at shorter than 1ms. This indicates that a considerable mole fraction of larger than 6 spin systems is present in mineral single crystal FAP. Thus, since the sum of the mole fraction of defect densities larger than 50% is greater than 90%, we can presume that the defect density is less than 50%. The contribution of larger than 6 spin systems to the MO dynamics hinders the reliable measurement of the defect density from comparisons between calculated and experimental plots. In spite of the existence of a variety of spin system in single crystal fluorapatite, the oscillatory behavior of MO intensities of single crystal FAP in fig. "-27 indicates that the fluorine atoms in chains are interrupted by defects. 6. Conclusions We have shown that multiple-quantum dynamics in an essentially one- dimensional distribution of spins leads to an initial linear dependence of effective size N upon preparation time, in qualitative agreement with predictions of the in- cremental shell model.36 Further work is needed to determine whether the lack of good quantitative agreement in this initial region arises from experimental consid- erations, or instead from limitations of the theoretical model, and whether further refinements of the model may improve its predictive accuracy. The MO growth exponents also distinguish the different dimensionalities of spin systems and the presence of clustering. The contribution of the two different dimensionalities to the growth of the MO coherences is visually illustrated by the MO growth exponent plot. Two growth exponents of HAP-M agree well reasonably with those of the 150 ‘ I 0.8 ‘ . he; 5 0.6 ' I g 8 1 g 0 § 0.4 . I c ‘ A A a 0.2 - 9 a a . O 8 . : I I: 00 LL 9 - . . t =. 0.0 0.2 0.4 0.6 0.8 Defect density figure "-35. Mole fraction of spins in various run numbers and sum of the mole fraction with different defect densities using percolation theory." Run number = 1 (open circle), run number =- 2 (open triangle), run number = 3 (open square), run number a 4 (cross in open square), run number a 5 (cross), and sum of 1 - 5 spins (black circle). 151 1.0 Cl 4 on b) $0 0.8 ‘ Cl DUO 13 1 A D D , DAAD U UAAA 0.6 D M AA 0 A 0.4 1 El AAA GAGA 83 26 A , A O 4‘ O @O O a a « is g l 'S 00 - r . . r . - ' - B . -% 1.0 g a) D b 5 0.8 1 D an a We a 0.6 . A AA DA A AARm DA 0 O A 0.4 mo 0 0.2 if? 0 0 0 o O 0.0 3 ' ' ' r r v r t v , - 0 200 400 600 800 1000 1200 Preparationtlfneols) j figure "-36. Comparison between experimental 20 intensities (open circle) of single crystal FAP and calculated 20 intensifies for two assumed defect densities [30% (open circle) and 50% (cpen square)] at two different orientations. a) 62° orientation with respect to the external magnetic field; b) 90° orientation with respect to the external magnetic field. 152 incremental shell model. Linear behavior of the effective sizes vs. preparation time has previously been observed generally only for finite clusters, but the slope at longer preparation times in these cases is observed to level off,57 in contrast to our results. It is interesting to note that the hopping model also predicts a linear dependence for clusters of 21 spins, but it cannot treat infinite spin systems, and has not predicted MO dynamics for different dimensionalities. It is feasible to model the one-dimensional chains due to their simple configu- ration. 1H MQ—NMR is sensitive to defects in one-dimensional apatite chains, and the defect density can be estimated by using the one-dimensional cluster model. The comparison of 4Q/2Q intensities between HAP-M and HAP-N for given prepa- ration times directly shows the presence of vacancies in the non-stoichiometric hydroxyapatite sample. The defect density of HAP-N obtained using MQ NMR is in good agreement with the 8% defect density obtained from 1H MAS-NMR;68 however, the calculated 4Q/ZO intensities based on the defect densities of the FOHAP samples deviate considerably from MQ experimental 40120 intensities. This discrepancy may arise from three possible sources; additional defects in the hydroxyl chains other than fluorine atoms (i. e. vacancies), non-random fluorine substitutions, and/or off-resonance effects caused by dipolar coupling of the hy- droxyl protons to 19F nuclei. Partially deuterated hydroxyapatite could be used to test the 1-D cluster model wi’d'l knowledge of the hydroxyl content, since a deuterium randomly substitutes for a proton and the dipolar interaction between a deuterium and a proton is small. The first explanation could be tested by a detailed chemical analysis of the hydroxyl content of the crystals, which is quite difficult to perform. Since the dipolar interaction in single crystal samples can be adjusted by “..- 153 field, we hope that 19F MO NMR of single crystal fluorapatite could be used to separate the MO dynamics of different dimensionalities. For example, if the orientation of intra-chain fluoride ions is fixed at 54.7° (magic angle), the non-zero dipolar coupling of inter-chain fluoride ions would create MQ coherence in a 2-D or fashion. Our 19F MQ-NMR results for a single crystal of mineral FAP show an oscillatory behavior of the MO peaks that does not permit measurement of the effective size at each preparation time. A less deficient synthetic single crystal sample might give information about dimensionality in more detail. Since the orientation of individual crystallites can be selected by relating them to a position in the CSA powder pattern, MQ-NMR can be performed on a specific orientation by selectively saturating all but a single frequency in the pattern.79 Selecting the orientation in a powdered sample thus also permits the relative magnitude of the . intra- and inter-chain dipolar couplings to be “tuned” at will. l! i”: “m5 11'- 55‘...” .\ VA . 154 7. References 1. W. A. Anderson, Phys. Rev. 104, 850 (1956) 2. J. I. Kaplan and S. Meiboom, Phys. Rev. 109, 499 (1957) 3. S. Yatsive, Phys. Rev. 113, 1522 (1959). 4. W. A. Anderson, R. Freeman, and C. A. Reilly, J. Chem. Phys. 39, 1518 (1963). 5. K. A. McLauchlan and D. H. Whiffen, Proc. Chem. 800., 144 (1962). 6. A. D. Cohen and D. H. Whiffen, Mal. Physics, 7, 449 (1964). 7. J. I. Musher, J. Chem. Phys. 40, 983 (1964). 8. M. L. Martin, G. J. Martin, and R. Couffignol, J. Chem. Phys. 49, 1985 (1968). 9. K. M. Worvill, J. Magn. Reson. 18, 217 (1975). 10. P. Bucci, M. Martinelli, and S. Santucci, J. Chem. Phys. 52, 4041 (1970). 11. P. Bucci, M. Martinelli, and S. Santucci, J. Chem. Phys. 53, 4524 (1970). 12. G. Lindblom, H. Wennerstroem, and B. Lindman, J. Mag. Reson. 23, 177 (1976) 13. R. B. Creel, E. D. von Meerwall, and R. G. Barnes, Chem Phys. Lett. 49, 501 (1977). 14. D. Dubber, K. Deorr, H. Ackmann, F. Fujara, H. Grupp, M. Grupp, P. Heitjans, A. Keoblein, and H. J. Steockmann, Z. Physik A, 282, 243 (1977). 15. R. R. Ernst and W. A. Anderson, Rev. Sci. Instr. 37, 93 (1966). 16. H. Hatanaka, T. Terao, and T. Hashi, J. Phys. Soc. Jpn. 39, 835 (1975). 17. W. P. Aue, E. Bartholdi, and R. R. Ernst, J. Chem. Phys. 64, 2229 (1976) 18. G. Bodenhausen, Prog. NMR Spectrosc. 14, 137 (1981). 19. U. Piantini, O. W. Sorenson, and R. R. Ernst, J. Am. Chem. Soc. 104, 6800 (1982). 155 20. A. Bax, R. Freeman, and S. P. Kempsell, J. Am. Chem. Soc. 102, 4849 (1980). 21. A. Bax, R. Freeman, and S. P. Kempsell, J. Magn. Reson. 41, 349 (1980). 22. A. Bax, R. Freeman, T. A. Frenkiel, and M. H. Levitt, J. Magn. Reson. 43, 478 (1981) 23. A. Bax, R. Freeman, and T. A. Frenkiel, J. Am. Chem. Soc. 103, 2102 (1981). 24. L. Braunschweiller, G. Bodenhausen, and R. R. Ernst, Mal. Phys. 48, 535 (1983). 25. M. H. Levitt and R. R. Ernst, Chem. Phys. Left. 100, 109 (1983). 26. W. S. Warren, D. P. Weitekamp, and A. Pines, J. Chem. Phys. 73, 2084 (1080). 27. W. S. Warren and A Pines, J. Chem. Phys. 74, 2808 (1981). 28. S. Sinton, D. B. Zax, J. B. Murdoch, and A. Pines, Mal. Phys. 53, 333 (1984). 29. W. S. Warren, S. Sinton, D. P. Weitekamp, and A. Pines, Phys. Rev. Left. 43, 1792 (1979). 30. Y. S. Yen and A. Pines, J. Chem. Phys. 66, 5624 (1983). 31. J. Baum, M. Munowitz, A. N. Garroway, and A. Pines, J. Chem. Phys. 83, 2015 (1985). 32. A. N. Garroway, J. Baum, M. Munowitz, A. Pines, J. Magn. Reson. 60, 337 (1984). 33. J. Baum, K. K. Gleason, A. Pine, A. N. Garroway, and J. A. Reimer, Phys. Rev. Left. 56, 1377 (1986). 34. K. K. Gleason, M. A. Petn'ch, and J. A. Reimer, Phys. Rev. 8.36, 3259 (1987). 156 35. M. A. Petrich, K. K. Gleason, and J. A. Reimer, Phys. Rev. 8.36, 9722 (1987). 36. D. H. Levy and K. K. Gleason, J. Phys. Chem. 96, 1992 (1992). 37. G. Drobny, A. Pines, S. Sinton, D. P. Weitekamp, and D. Wemmer, Faraday Symp. Chem. Soc. 13, 49 (1979). 38. R. Ryoo, S.-B. Liu, L. C. de Menorval, K. Takegoshi, B. Chmelka, M. Trecoske, and A. Pines, J. Phys. Chem. 91, 6575 (1987). 39. B. Chmelka, J. G. Pearson, S.-B. Liu, R. Ryoo, L. C. de Menorval, and A. Pines, J. Phys. Chem. 95, 303 (1991). 40. D. N. Shykind, J. Baum, S.-B. Liu, A. Pines, and A. N. Garroway, J. Magn. Reson. 76, 149 (1988). 41. S. Sinton and A. Pines, Chem. Phys. Left. 76, 263 (1980). 42. M. Munowitz and A. Pines, Science 233, 525 (1986). 43. W. V. Gerasimowicz, A. N. Garroway, and J. B. Miller, J. Am. Chem. Soc. 1 12, 3726 (1990). 44. B. E. Scruggs and K. K. Gleason, J. Magn. Reson. 99, 149 (1992). 45. B. E. Scruggs and K. K. Gleason, Macromolecules, 25(7), 1864 (1992). 46. M. Gold man, “Quantum Description of High-Resolution NMR in Liquids” (Oxford University Press, New York, 1988). 47. C. P. Slichter, “Principles of Magnetic Resonance”, 3rd ed. (Springer-Verlag, New-York, 1989). 48. R. R. Ernest, G. Bodenhausen and A. Wokaun, “Principles of Nuclear Magnetic Resonance in One and Two Dimensions” (Oxford University Press, New York, 1987). 1,, 49. 50. 51 . 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 65. 66. 67. 68. 157 D. P. Weitekamp, Adv. Magn. Reson. 11, 111 (1983). M. Munowitz, Mol. Phys. 71, 959 (1990). B. E. Scruggs and K. K. Gleason, Chem. Phys. 166, 367 (1992). M. Munowitz and A. Pines, Adv. Chem. Phys. 66, 1 (1987). S. Lacelle, Advances in Magnetic and Optical Resonance 16, 173 (1991 ). A. Wokaun and R. R. Ernst, Mol. Phys. 36, 317 (1978). J. Baum and A. Pines, J. Am. Chem. Soc. 108, 7447 (1986). J. B. Murdoch, W. S. Warren, D. P. Weitekamp and A. Pines, J. Mag. Reson. 60, 205 (1984). M. Munowitz, A. Pines, and M. Mehring, J. Chem. Phys. 86, 3172 (1987). M. I. Kay, R. A. Young and A. S. Posner, Nature (London), 204, 1050 (1964). J. C. Elliott, P. E. Mackie, and R. A. Young, Science, 180, 1055 (1980). A. Abragam, "Principles of Nuclear Magnetism" Oxford Univ. Press, Oxford (1961). M. Munowitz, “Coherence and NMR’ Wiley, New York (1988). J. von Neumann, “Mathematical Foundations of Quantum Mechanics”, Princeton University Press. Princeton, NJ (1955). U. Haeberlen, "High Resolution NMR in Solids: Selective Averaging", Adv. Magn. Reson., Supplement 1, . Academic Press, New York (1976). U. Haeberlen and J. S. Waugh, Phys. Rev. 175, 453 (1968). J.L. Carolan, Chem. Phys. Left. 12 (1971) 389. F. S. de Bouregas and J. S. Waugh, J. Magn. Reson. 96, 280 (1992) BC. Fowler, Inorg. Chem. 13 (1974) 207. J. Arends, J. Christofferson, M. R. Christofferson, H. Eckert, B. O. Fowler, J. C. Heughebaert, G. H. Nancollas, J. P. Yesinowski and S. J. Zawacki, “"3 158 J. Cryst. Growth 84, 515 (1987). 69. J. P. Yesinowski and H. Eckert, J. Am. Chem. Soc. 109, 6274 (1987). 70. EC. Moreno, M. Kresak, and RT Zaradnik, Caries Res. Supp/.1 142 (1977) 71. JP Yesinowski and M.J. Mobley, J. Am. Chem. Soc.105 6191 (1983); JP Yesinowski, R. Wolfgang, and M.J.. Mobley, in Adsorption on and Surface Chemistry of Hydroxyapatite, Mlsra, D. N., Ed., Plenum: New York, pp151-175. 72. R. W. C. Wyckoff, Crystal Structures (lnterscienoe New York), 6, part 1,386 (1963). 73. E. R. Andrew, J. Chem. Phys. 18, 607 (1950). 74. L. B. Schreiber and R. W. Vaughan, Chem. Phys. Left., 28, 586 (1974). 75. K. K. Gleason, Concepts in Magn. Reson. Appeared. . 76. W. Van der Lugt, D. l. M. Knottnerus, and W. G. Perdok, Acta Crysallogr., Sect. 8: Crystallogr. Cryst. Chem. B27, 1509 (1971). 77. A. M. Vakhrameev, S. P. Gaguda, and R. G. Knubovets, J. Stmct. Chem. (Engl. Transl.), 19, 256 (1978). 78. D. Stauffer and A. Aharony, “Introduction to Percolation Theory” Taylor & Francis, London, Washington, DC (1992). 79. L. B. Moran and J. P. Yesinowski, Chem. Phys. Left., To be submitted. -D‘i“;; hffi.m1 I 159 APPENDIX A Pulse program for even-order selective phase-incremented or TPPI multiple- quantum pulse sequence for solids for Varian VXR spectrometer. This program is courtesy of T. Barbara of Varian Associate. /*Pulse sequence: G_INMQ_LOCK*/ [*Pseudo l—D even order MQ generation with a detection using spin locking sequence. Phase cycle to remove odd order quanta by alternating excitation phase between x and -x and adding memory.*/ /* VARIABLES: del: delay between 8 pulses ps delp: delay between 8 pulses, 2*del-l-pw ps mloop: the number of loop (basic cycle time*mloop = preparation time) shift: the angle phase incrementation (determine the maximum detectable MQ order) tlinc: the time to increment the phase d3: the time between mixing and detection period to eliminate the remaining transverse magnetization d4: the evolution time*/ #include static int table 1[4] = static int tab182[4] = static. int table3[4] = static int table4[4] = static int tableS[4] = static int tab1e6[4] = static int table7[4] = 9 9 1 9 9 C 9 9 {0220 {0022 {2002 {1,1,3.3 {3311 {0202 {1313 O }; }; . , }; } i 9 9 9 9 9 9 pulse sequenceO { double del, delp, ddel, ddelp, mloop, shift, d3, d4, tlinc: char trig[MAXSTR]; extem double getvalO; del = getva1(“del”); delp = gctva1(“delp”); 160 mloop = getval(“mloop”); shift = getval(“shift”); d3 = getval(“d3”); d4 = getval(“d4”); tlinc = getval(“tlinc”); getstr(“trig”,trig); ddel = del - rofl - rof2; ddelp = del - rofl - rof2; settable(tl, 4, tablel); settable(t2, 4, table2); settable(t3, 4, table3); settable(t4, 4, table4); settable(tS, 4, tableS); settable(t6, 4, tableé); settable(t7, 4, table7); setreceiver(t7); stepsize(shift,TODEV); initval(mloop.v 1); initva1( np/2.0,v7); initdelay(tlinc,DELAY5); assign(zero,v 10); loop(v7,v9); if(trig[0 =’y’) { xgate(l .0); } xmtrphase(v10); delay(d1); delay(rofl); incr(v10); rcvroffO; starthardloop(vl); delay(ddel); rgpulse(pw,tl,rofl,rof2); f" x -x -x x */ delay(ddelp); rgpulse(pw,tl ,rofl ,rot‘Z); delay(ddel); rgpulse(pw,tl,rofl ,rof2); delay(ddelp); 161 rgpulse(pw,tl,rof1,rof2); delay(ddel); rgpulse(pw,t3,rof1,rof2); /* -x x x -x */ delay(ddelp); rgpulse(pw,13,rofl ,rof2); delay(ddel); rgpulse(pw,t3,rof1,rof2); delay(ddelp); rgpulse(pw.t3.rofl.rof2); endharleOPO; xmuphase(zero); /* reset small angle shift to zero */ delay(d4); incdelay(v10,DELAY5); /* generate t1 delay If tlinc=0, TPPI*/ /*otherwise, phase incremented method*/ stanhardloop(v 1); delay(ddel); rgpulsc(pw,t4,rofl,rof2); f“ y y -y -y */ delay(ddelp); rgpulse(pw,t4,rofl ,ron); delay(ddel); rgpulse(pw,t4,rof1,rof2); delaflddelp); rgpulse(pw,t4,rof1,rof2); delay(ddel); rgpulse(pw,t5,rof1,rot?); /* -y -y y y */ delay(ddelp); rgpulse(pw,t5,rof1,roa); delay(ddel); rgpulse(pw,15 ,rof l ,rof2); delay(ddelp); rgpulse(pw,t5,rof1,rof2); endhardloopO; rcvron(); delay(d3); /* delay to allow decay of transverse coherences */ rgpulse(pw, t6, rofl, rof2); I“ x -x x -x */ rgpulse(p1, t7, rofl, rof2); /* y -y y -y spin loclcing*/ delay(alfa); acquirc(2.0,l/sw); endloop(v9); I FL" f5 162 APPENDIX B Selection rule of 1-spin/2-quantum average Hamiltonian Hy, in hopping model If the spin angular momentum operator 1,, is expressed for a spin with index 3,- by a ket jejaj). the Liouville state that is formed is a product of K single-spin operators that can be written by the abbreviated notation |Knp)=lslal ............ 850k) ' ‘ (A-1) where n = Z or". (A'2) j=l,K The scalar product between two Liouville state vectors is orthornormal and is de- fined as (AlB) = Tr{A+B} (A-3) where A+ is the Hermitian adjoint of A. Since the superoperator 7:1 is defined by 710.4) = |[’H, .41), (A-4) matrix elements involving the superoperator are given by (AWE) -_- (Al [71,19]) -_- Tr{A+[H,B]}. (A-5) The selection rules in the Liouville space from the Hamiltonian in Eq. lI-47 are calculated using Eq. A-5 and the commutation relations [I,,-, I,,] = A6,,- (A-6a) and [I+i1I—j1 = 21260 (A-6b) Since the selection rule of the average dipolar Hamiltonian from the pulse se- quence in Fig. ll-4 can be evaluated using the numerator of Eq. Il-47, we can rewrite Eq. A-5 as (KnplfiDIK’n’p’) = (1(an [120,K’n'p’]). (A-7) After K ' = 1 and n' = 0 (11,) is chosen as an initial condition, the square bracket 163 part of Eq. A-7 can be written using only the spin operators [[+1I+2 + Lil—2, [12] = [[+1I+2,1'1z] + [1—11-2, I121 (A-8) using a commutator algebra [A, B + C] = (A, B] + [A, C]. (A-9) Since 1+2 and 1.2 commute with [1,, we can rewrite Eq. A-8 using the commuta- tion relations in Eq. A-6 as = I+2[I+1.1121+I—2lI—1ll'12] = -I+2I+1-I—2I—1 (Pl-10) [.4. BC] = [A, B]C + B[A, C]. (A-11) The values of K and n of the first term in Eq. A-10 are 2 and 2, and those of the second term in Eq. A-10 are 2 and -2 respectively. Therefore, the solution of Eq. A-7 is . = (Knpl - 1+2I+1 - 1.2111) = (K711018138)+- (A42) ' Since the scalar product of the Liouville state is orthonormal, Eq. A-12 is non- zero at K-2, n=2 and K-2, n--2. When It" a 2 and n' =2 ( from the 1,21“ term), [HD,K'n'p'] can be also written using the spin operators [1+1I+2 + 1.11-2,I+2I+1] = [1+1I+2, 1+2I+11 + [I—tI—z, 1+2I+1] (A-13) Since the first term [I+II+2,I+2I+1] commutes, we solve for the second term [I_1I_2,I+2I+1] using Eq. A-11 = [+211—11—2J+11 1' 11-11—2. I+211+1 = -I+1I—1122 + 1121—21” (A44) Therefore, only when K=3 and n=0 is (Knpl’lez, 2, p) non-zero. The detailed calculation of the selection rule for K ' = 2 and n' =- -2 is omitted since it is similar to the calculation of the selection rule for It" = 2 and n' = 2. Only when K = 3 and n = 0 is (Knpl’lefz,—2, p) non-zero too. 164 We will calculate the general solution of the selection rule using Eq. A-7. The square bracket in Eq. A-7 can be represented by [Hal K 'n'P'] = [.2 (1+il'+j + I_,-I_,-), K'nlp'] (A'1 5) 1)] We can rewrite Eq. A-15 using Eq. A-9 and A-11 as = 2 1+,- [I+j,K'n'p'] + Z [LI-itK'nIP']I-l-I i>j t'>j + XI..- [I_,-,K'n'p’] + 2 [I_,,K’n'p']I_,. (A-16) The solutions of): [1+], K 'n'p'] and :2: [I+,-, K 'n'p'] do not change the value of K ' but change 12' tr]; n' + 1 due to the commutation relations (Eq. A-6a and A-6b). The lowering operators in the square brackets of Eq. (A-16) do not change the value of K' but decrease n' to n' — 1. Since the operators outside the square brackets in Eq. A-16 increase K ' to K ' + 1, and change n' to n' :t 1 according to (Knplflle'n'p'> is non-zero when K is equal to K ' + 1 and n is equal to n' j: 2. When we solve the Eq. A-7 as(KnijD|K'n'p') = ([Knp,HD]jK'n'p'), the non-zero condition of (Knp|Hp|K'n'p'> is that K is equal to K ' — 1 and n is equal to n' :l: 2. Therefore, the selection rule of the average Hamiltonian 71,, in the Liouville spaceisK=K'i1andn=n'i2. 165 APPENDIX C C programs for calculation of the ratio of 40/20 intensities for a 1-D chain with different defect densities. These programs calculate the ratio of the 4Q/2Q intensities of defect containing 1D chains for various preparation times using the statistical model, combined with percolation theory expressions that provide the mole fraction of spins in a run number. /* Calculation of 4Q/2Q intensity using the statistical model and the percolation theory*/ #include #include #include double x,y; double exv().sqfl().exr>2().pow0; mainO { int i,j,n,l,m,p; float k,s,h,sum2,pp,sump2,sum2p,sum4,sump4,sum4p,rat2,rat4; char filename[20]; FILE *fp; printf("Enter filename: \n"); scanf("%s",filenamc); fp=fopen(filename."wr"); pfintt‘C'What is the ratio of the defect’?\n"); fprintf(fp,"What is the ratio of the defect. "); scanf("%f",&k); fprintf(fp,"%f\n", k); printf("What is the maximum run number. "); fprintf(fp,"What is the maximum run number?\n"); scanf("%d",&l); 166 fprintf(fp,"%d\n", 1); printf("What is the efi'ective size?\n"); fprintf(fp,"What is the effective size?\n"); scanf("%f",&h); fprintf(fp,"%f\n". h); s=1.0-k; sum2=0.0; sum4=0.0; sum2p=0.0; sum4p=0.0; [*The first part of Eq. 11-82 for 2-Q intensity. The decimal number (efiective size from Gaussian function) cannot be used to calculate a factorial so that the round up values are used.*/ if(h<=6.0) { for(i=2;i<=l;i++) { x=(doub1e)i; if(x<=h) {s}um2+=x*pow(k,2.0)*pow(s,x)*fac(2*i)/(fac(i-2)*fac(i+2))/pow(4.0,x); else { P=round(h); PP=(float)p; sum2 +=x*pow(k,2.0)*pow(s,x)*fac(2*p)/(fac(p-2)*fac(p+2))/pow(4.0,pp); 1 1 [*The first part of Eq. II-82 for 4-Q intensity. The decimal number (efi‘ective size from Gaussian function) cannot be used to calculate a factorial so that the round up values are used.*/ for(ifl;i<=l;i++) { 167 x=(double)i; if(x<=h) { sum4 +=x*pow(k,2.0)*pow(s,x)*fac(2*i)/(fac(i-4)*fac(i+4))/pow(4.0,x); 1 else 1 p=round(h); pp=(float)p; sum4 +=x*pow(k,2.0)*pow(s,x)*fac(2*p)/(fac(p-4)*fac(p+4))/pow(4.0,pp); } } 1 /*The second part of Eq. II-82 for 2Q and 4Q intensity. */ else I for(i=2;i<=6;i++) { x=(double)i; sum2p +=x*pow(k,2.0)*pow(s,x)*fac(2*i)/(fac(i-2)*fac(i+2))/pow(4.0,x); } for(i=4;i<=6;i-H-) { x=(double)i; sum4p +=x*pow(k,2.0)*pow(s,x)*fac(2*i)/(fac(i-4)*fac(i+4))/pow(4.0,x); } . for(i=7;i<=1;i-H-) { x=(double)i; if (x<=h) { sum2p +=x*pow(k,2.0)*pow(s,x)l(sqrt(x)*1.77)*exp(-4.0/x); sum4p +=x*pow(k,2.0)*pow(s,x)/(sqrt(x)* 1 .77)*exp(-16.0/x); 1 else 1 sum2p +=x*pow(k,2.0)*pow(s,x)/(sqrt(h)* 1 .77)*exp(-4.0/h); 168 sum4p +=x*pow(k,2.0)*p0W(8.x)/(Sq1't(h)*1-77)*6XP(' 16W“); } 1 } printf("The ratio of each multiple quantum signal\n"); fprintf(fp,"The ratio of each multiple quantum signal\n"); sump2=(sum2+sum2p)/(sum2+sum2p); sump4=(sum4—I-sum4p)/(sum2+sum2p); printf("2quantum => %f, 4quantum => %f\n",sump2,sump4); fprintf(fp,"2quantum ==> %f, 4quantum => %t\n",sump2,sump4); fclose(fp); } [*Calculation of factorial’V fac(m) int m; { int i, fac; fac=l; for(i=l;i<=m;i-H~) fac=fac*i; rettn'n(fac); } /* Calculation of round up*/ round(x) ' float x; 1 int ix; float fx, gx; 8X=X; ix=(int)gX; fx=(float)ix; gx-=fx; if(gx>=05) ix+=1; returnfiX): I “‘illillllillili“