dk1719 ch10

33
10 Solid-State Nuclear Magnetic Resonance Spectroscopy David E. Bugay SSCI, Inc., West Lafayette, Indiana I. INTRODUCTION Nuclear magnetic resonance (NMR) spectroscopy has become a fundamental technique for pharmaceutical analysis (1–8). Typically, when one thinks of NMR spectroscopy, solution-phase studies, such as structure elucidation, immediately come to mind (1). Within a drug discovery setting, structure elucidation is the one of the primary uses of NMR. It is ironic to think that although approximately 80% of all pharmaceutical products are solid-state formulations, solid-state NMR is relatively new to pharmaceutical analysis. Based on the myriad of uses of NMR spectroscopy as applied to pharmaceutical analysis, this chapter focuses on the use of solid-state NMR for the analysis of pharmaceuticals. Since the advent of solid-state NMR in the 1970s, an increasing number of applications of solid-state NMR to the study of pharmaceuticals have been published. Probably the most popular use of solid-state NMR has been in the study of polymorphism (9,10). Traditionally, polymorphism has been investi- gated by X-ray diffraction and thermal analysis techniques. Various solid-state NMR methodologies and pulse sequences have shed new light on the understand- ing of polymorphism. An extensive review of scientific advances within this as- pect of drug development will be reviewed in this work. Additionally, solid-state NMR has been utilized for studying drug–excipient interactions, drug–membrane interactions, chirality, structural characterization, and molecular mobility—a key aspect of solid-state reactions. This chapter provides a basis of theory for solid-state NMR and a brief review of experimental aspects. In the discussion section, the various applications Copyright 2002 by Marcel Dekker. All Rights Reserved.

Upload: sumant-saini

Post on 10-May-2015

310 views

Category:

Documents


23 download

TRANSCRIPT

Page 1: Dk1719 ch10

10Solid-State Nuclear MagneticResonance Spectroscopy

David E. BugaySSCI, Inc., West Lafayette, Indiana

I. INTRODUCTION

Nuclear magnetic resonance (NMR) spectroscopy has become a fundamentaltechnique for pharmaceutical analysis (1–8). Typically, when one thinks of NMRspectroscopy, solution-phase studies, such as structure elucidation, immediatelycome to mind (1). Within a drug discovery setting, structure elucidation is theone of the primary uses of NMR. It is ironic to think that although approximately80% of all pharmaceutical products are solid-state formulations, solid-state NMRis relatively new to pharmaceutical analysis. Based on the myriad of uses of NMRspectroscopy as applied to pharmaceutical analysis, this chapter focuses on theuse of solid-state NMR for the analysis of pharmaceuticals.

Since the advent of solid-state NMR in the 1970s, an increasing numberof applications of solid-state NMR to the study of pharmaceuticals have beenpublished. Probably the most popular use of solid-state NMR has been in thestudy of polymorphism (9,10). Traditionally, polymorphism has been investi-gated by X-ray diffraction and thermal analysis techniques. Various solid-stateNMR methodologies and pulse sequences have shed new light on the understand-ing of polymorphism. An extensive review of scientific advances within this as-pect of drug development will be reviewed in this work. Additionally, solid-stateNMR has been utilized for studying drug–excipient interactions, drug–membraneinteractions, chirality, structural characterization, and molecular mobility—a keyaspect of solid-state reactions.

This chapter provides a basis of theory for solid-state NMR and a briefreview of experimental aspects. In the discussion section, the various applications

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 2: Dk1719 ch10

of solid-state NMR as applied to pharmaceuticals are outlined. Qualitative charac-terization techniques are summarized as well as quantitative methodology usedfor the study of the active pharmaceutical ingredient (API) or the final drugproduct.

II. SOLID-STATE NMR THEORY

Conventional utilization of solution-phase NMR data acquisition techniques onsolid samples yields broad, featureless spectra (Fig. 1A). The broad nature of thesignal is due primarily to dipolar interactions, which do not average out to zeroin the solid state, and chemical shift anisotropy (CSA), which again is a functionof the fact that our compound of interest is in the solid state. Before one describesthe two principal reasons for the broad, featureless spectra of solid-state materials,it is important to understand the main interactions that a nucleus with a magneticmoment experiences when situated within a magnetic field. In addition, manifes-tations of these interactions in the solid-state NMR spectrum need to be discussed.

A. Zeeman Interaction

The principal interaction that a nucleus experiences when placed in an externallyapplied magnetic field (B0, tesla), is the Zeeman interaction (HZ), which describesthe interaction between the magnetic moment of the nucleus and B0. The nuclearmagnetic moment (µ, ampere meter2) is proportional to the nuclear spin quantumnumber (I) and the magnetogyric ratio (γ, radian tesla�1 second�1), Eq. (1)(h � Planck’s constant). Thus, the Zeeman interaction occurs only with nucleiwhich possess a spin greater than zero, and yields 2I � 1 energy levels of

µ �γIh2π

(1)

HZ � �γ�B0 IZ (2)

separation ν0 � γB0/2π. Equation (2) describes the Hamiltonian, where ν0 is thecorresponding Larmor frequency (� � h/2π). The interaction is linear with theapplied magnetic field, thus giving the impetus to manufacture higher-magnetic-

Fig. 1 Solid-state 13C NMR spectra of didanosine (VIDEX-ddI) acquired by: (A) con-ventional solution-phase pulse techniques; (B) high-power proton decoupling only; (C)high-power proton decoupling and magic angle spinning (5 kHz); (D) high-power protondecoupling combined with magic angle spinning (MAS) and cross-polarization (CP). Ineach case, 512 scans were accumulated. (From Ref. 11.)

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 3: Dk1719 ch10

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 4: Dk1719 ch10

field spectrometers since a larger separation of energy levels leads to a greaterpopulation difference, a subsequent increase in the signal-to-noise ratio (S/N),and increased spectral resolution. Since the Zeeman interaction incorporates themagnetogyric ratio, a constant for each particular nucleus, the resonant frequencyfor each nucleus is different at a specific applied magnetic field (Table 1). Themagnitude of the Zeeman interaction for a 13C nucleus in a 5.87-T magnetic fieldis 62.86 MHz. Small perturbations to the Zeeman effect are produced by otherinteractions, such as dipole–dipole, quadrupolar, shielding, and spin–spin cou-pling. Typically, these small perturbations are less than 1% (�600 kHz) of theZeeman interactions, which range in frequency from 106 to 109 Hz.

B. Dipole-Dipole Interaction

Dipole–dipole interaction is the direct magnetic coupling of two nuclei throughspace. This interaction may involve two nuclei of equivalent spin or nonequiva-lent spin, and is dependent on the internuclear distance and dipolar couplingtensor. Additionally, the total interaction, labeled HD, is the summation of allpossible pairwise interactions (homo- and heteronuclear). It is important to notethat the interaction is dependent on the magnitude of the magnetic moments,which is reflected in the magnetogyric ratio, and on the angle (θ) that the in-ternuclear vector makes with B0. Therefore, this interaction is significant for spin-1/2 nuclei with large magnetic moments, such as 1H and 19F. Also, the interactiondecreases rapidly with increasing internuclear distance (r), which generally corre-sponds to contributions only from directly bonded and nearest-neighbor nuclei.Equation (3) describes the dipolar interaction for a pair of nonequivalent, isolatedspins I and S. Since the dipolar coupling tensor, D, contains a (1 � 3 cos2 θ)term, this interaction is dependent on the orientation of the molecule. In

H ISD �

γIγS�2

r 3I ⋅ D ⋅ S (3)

solution-phase studies where the molecules are tumbling rapidly, the (1–3cos2 θ) term is integrated over all angles of θ and subsequently disappears. Insolid-state NMR, the molecules are fixed with respect to B0, thus the (1–3 cos2

θ) term does not approach zero. This leads to broad resonances in the solid-stateNMR spectrum, since dipole–dipole interactions typically range from 0 to 105

Hz in magnitude.In the case of pharmaceutical solids which are dominated by carbon and

proton nuclei, the dipole–dipole interactions may be simplified. The carbon andproton nuclei may be perceived as ‘‘dilute’’ and ‘‘abundant’’ based on their isoto-pic natural abundance, respectively (Table 1). Homonuclear 13C–13C dipolar inter-

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 5: Dk1719 ch10

Table 1 NMR Frequency Listing

NMR frequencySensitivity (MHz) at a field (T) of

NaturalIsotope Spin abundance rel.a abs.b 2.35 5.87 11.74

1H 1/2 99.98 1.00 1.00 100.00 250.00 500.002H 1 1.5 � 10�2 9.65 � 10�3 1.45 � 10�6 15.35 38.38 76.7510B 3 19.58 1.99 � 10�2 3.90 � 10�3 10.75 26.87 53.7311B 3/2 80.42 0.17 0.13 32.08 80.21 160.4213C 1/2 1.11 1.59 � 10�2 1.76 � 10�4 25.14 62.86 125.7214N 1 99.63 1.01 � 10�3 1.01 � 10�3 7.22 18.06 36.1215N 1/2 0.37 1.04 � 10�3 3.85 � 10�6 10.13 25.33 50.6617O 5/2 3.7 � 10�2 2.91 � 10�2 1.08 � 10�5 13.56 33.89 67.7819F 1/2 100.00 0.83 0.83 94.08 235.19 470.3923Na 3/2 100.00 9.25 � 10�2 9.25 � 10�2 26.45 66.13 132.2625Mg 5/2 10.13 2.67 � 10�3 2.71 � 10�4 6.12 15.30 30.6027Al 5/2 100.00 0.21 0.21 26.06 65.14 130.2929Si 1/2 4.7 7.84 � 10�3 3.69 � 10�4 19.87 49.66 99.3331P 1/2 100.00 6.63 � 10�2 6.63 � 10�2 40.48 101.20 202.4033S 3/2 0.76 2.26 � 10�3 1.72 � 10�5 7.67 19.17 38.3535Cl 3/2 75.53 4.70 � 10�3 3.55 � 10�3 9.80 24.50 48.9937Cl 3/2 24.47 2.71 � 10�3 6.63 � 10�4 8.16 20.39 40.7839K 3/2 93.1 5.08 � 10�4 4.73 � 10�4 4.67 11.67 23.3341K 3/2 6.88 8.40 � 10�5 5.78 � 10�6 2.56 6.40 12.8143Ca 7/2 0.145 6.40 � 10�3 9.28 � 10�6 6.73 16.82 33.6467Zn 5/2 4.11 2.85 � 10�3 1.17 � 10�4 6.25 15.64 31.2779Br 3/2 50.54 7.86 � 10�2 3.97 � 10�2 25.05 62.63 125.2781Br 3/2 49.46 9.85 � 10�2 4.87 � 10�2 27.01 67.52 135.03127I 5/2 100.00 9.34 � 10�2 9.34 � 10�2 20.00 50.02 100.04195Pt 1/2 33.8 9.94 � 10�3 3.36 � 10�3 21.50 53.75 107.50

a Determined for an equal number of nuclei at a constant field.b Product of the relative sensitivity and natural abundance.Source: From Ref. 11.

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 6: Dk1719 ch10

actions essentially do not exist because of the low concentration of 13C nuclei(natural abundance of 1.1%). On the other hand, 1H–13C dipolar interactions con-tribute significantly to the broad resonances, but this heteronuclear interactionmay be removed through simple high-power proton decoupling fields, similar tosolution-phase techniques.

C. Chemical Shift Interaction

The three-dimensional magnetic shielding by the surrounding electrons is an ad-ditional interaction that the nucleus experiences in either the solution or the solidstate. This chemical shift interaction (HCS) is the most sensitive interaction tochanges in the immediate environment of the nucleus and provides the most diag-nostic information in a measured NMR spectrum. The effect originates from thesmall magnetic fields that are generated about the nucleus by currents inducedin orbital electrons by the applied field. These small perturbations upon the nu-cleus are reflected in a change in the magnetic field experienced by the nucleus.Therefore, the field at the nucleus is not equal to the externally applied field andhence the difference is the nuclear shielding, or chemical shift interaction [Eq.(4)]. It is important to note the orientation dependence of the shielding constant,σ, and the fact that shielding is proportional to the applied field, hence theneed for chemical shift reference materials such as tetramethylsilane.

HCS � γI �I ⋅ σ ⋅ B0 (4)

Solution-state NMR spectra yield ‘‘average’’ chemical shift values whichare characteristic of the magnetic environment for a particular nucleus. The aver-age signal is due to the isotropic motion of the molecules in solution. In otherwords, B0 ‘‘sees’’ an average orientation of a specific nucleus. For solid-stateNMR, the chemical shift value is also characteristic of the magnetic environmentof a nucleus, but normally, the molecules are not free to move. It must be keptin mind that the shielding will be characteristic of the nucleus in a particularorientation of the molecule with respect to B0. Therefore, a specific functionalgroup oriented perpendicular to the magnetic field will give a sharp signal charac-teristic of this particular orientation (Fig. 2A). Analogously, if the functionalityis orientated parallel to B0, then a sharp signal characteristic of that orientationwill be observed (Fig. 2B). For most polycrystalline pharmaceutical samples, arandom distribution of all orientations of the molecule will exist. This distributionproduces all possible orientations and is thus observed as a very broad NMRsignal (Fig. 2C). The magnitude of the chemical shift anisotropy is typically be-tween 0 and 105 Hz.

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 7: Dk1719 ch10

Fig. 2 Schematic representation of the 13C NMR signal of a single crystal containingthe functional group A-B, orientated (A) perpendicular to the applied field, and (B) parallelto the applied field. The lineshape in (C) represents the NMR signal of a polycrystallinesample with a random distribution of orientations yielding the chemical shift anisotropypattern displayed. (From Ref. 11.)

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 8: Dk1719 ch10

D. Spin–Spin Couplings and Quadrupolar Interactions

Two additional interactions experienced by the nucleus in the solid state are spin–spin couplings to other nuclei and quadrupolar interactions, which involve nucleiof spin greater than 1/2. Spin–spin (HSC), or J coupling [Eq. (5)], originates fromindirect coupling between two spins by means of their electronic surroundingsand are several orders of magnitude smaller (possibly 0–104 Hz, typically onlyseveral kilohertz) than dipole interactions. Although J coupling interactions occurin both the solid and the solution state, these couplings are typically used insolution-phase work for conformational analysis (12). Quadrupolar interactions(HQ) arise from dipolar coupling of quadrupolar nuclei, which display a non-

HSC � I ⋅ J ⋅ S (5)

spherically symmetrical field gradient, with nearby spin-1/2 nuclei [Eq. (6)]. Themagnitude of the interaction is dependent on the relative magnitudes of thequadrupolar and Zeeman interactions and may completely dominate the spec-trum (0–109 Hz), but since pharmaceutical compounds are primarily hydro-carbons, quadrupolar interactions typically do not interfere.

HQ � I ⋅ Q ⋅ I (6)

Summarizing the interactions, the isotropic motions of molecules in thesolution state yield a discrete average value for the scalar spin–spin coupling andchemical shift interactions. For the dipolar and quadrupolar terms, the averageobtained is zero, and the interaction is not observed in solution-phase studies. Insharp contrast, interactions in the solid state are orientation-dependent, subse-quently producing a more complicated spectrum, but one that contains muchmore information. In order to yield highly resolved, ‘‘solution-like’’ spectra ofsolids, a combination of three techniques is used: dipolar decoupling (13), magic-angle spinning (14,15), and cross-polarization (16). The remainder of this sectiondiscusses the three techniques and their subsequent effect on measured spectraof pharmaceutical compounds which are dominated by carbon nuclei.

E. Dipolar Decoupling

Discussed earlier was the fact that high-powered proton decoupling fields caneliminate the heteronuclear 1H–13C dipolar interactions that may dominate a solid-state NMR spectrum. The concept of decoupling is familiar to the solution-phaseNMR spectroscopist, but needs to be expanded for solid-state NMR studies. Insolution-phase studies, the decoupling eliminates the scalar spin–spin coupling,not the dipole–dipole interactions (this averages out to zero due to isotropic mo-tions of the molecules). Irradiation of the sample at the resonant frequency of

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 9: Dk1719 ch10

the nucleus to be decoupled (B2 field) causes the z component of the spins to fliprapidly compared to the interaction one wishes to eliminate. Scalar interactionsusually require 10 W of decoupling power or less. In pharmaceutical solids work,decoupling is used primarily to remove the heteronuclear dipolar interactionsbetween protons and carbons. The magnitude of the dipolar interaction (�50kHz) usually requires decoupling fields of 100 W and subsequently removes bothscalar and dipole interactions. Even with the use of high-power decoupling, broadresonances still remain, principally due to chemical shift anisotropy (Fig. 1B).

F. Magic-Angle Spinning

Molecules in the solid state are in fixed orientations with respect to the magneticfield. This produces chemical-shift anisotropic powder patterns for each carbonatom, since all orientations are possible (Fig. 2). It was shown as early as 1959that rapid sample rotation of solids narrowed dipolar-broadened signals (14). Anumber of years later, it was recognized that spinning could remove broadeningcaused by CSA yet retain the isotropic chemical shift (15).

The concept of magic-angle spinning arises from the understanding of theshielding constant, σ [Eq. 4]. This constant is a tensor quantity and, thus, can berelated to three principal axes:

σzz � 3

i�1

σ i cos2 θ i (7)

where σ i is the shielding at the nucleus when B0 aligns along the ith principalaxis, and θ i is the angle this axis makes with B0. Under conditions of mechanicalspinning, this relationship becomes time dependent and a (3 cos2 θ � 1) termarises. By spinning the sample at the so-called magic-angle of 54.7°, or 54°44′,this term becomes zero and thus removes the spectral broadening due to CSA(Fig. 1C), provided the sample rotation (kHz) is greater than the magnitude ofthe CSA (kHz).

CSA may range from 0 to 20 kHz, so our spin rates must exceed this valueor spinning sidebands are observed. Figure 3A displays the solid-state 31P-NMRspectrum of fosinopril sodium (Monopril), a novel ACE inhibitor (17), acquiredunder proton decoupling and static spinning conditions. At a relatively low spinrate of 2.5 kHz, broadening due to CSA is removed and the center band andsidebands appear (Fig. 3B). As the sample is spun at higher rates (Figs. 3C and3D), the sidebands become less intense and move out from the center band. Evenat a spin rate of 6 kHz (maximum spin rate for the probe/instrument used), theCSA is not completely removed (Fig. 3E). Although fast enough spin rates maynot be achieved to remove CSA totally for specific compounds, slower than opti-mal rates will still narrow the resonances. Spinning sidebands, at multiples of

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 10: Dk1719 ch10

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 11: Dk1719 ch10

the spin rate, will complicate the spectrum, but they can be easily identified byvarying the spin rate and observing which signals change in frequency. In addi-tion, specific pulse sequences may be used to minimize the spinning sidebands(18). While increasing the magnetic field strength increases the signal-to-noiseratio (Zeeman interaction), it also increases the CSA, since this interaction isfield-dependent [Eq. (4)]. Utilizing today’s high-field spectrometers (4.7 T),spinning sidebands may exist but can be identified and used to gain additionalinformation or be potentially eliminated if necessary.

The techniques of magic-angle spinning and heteronuclear dipolar decou-pling produce solid-state NMR spectra which approach the linewidths and ap-pearance of solution-phase NMR spectra. Unfortunately, there is an inherent lackof sensitivity in the general NMR experiment due to the nearly equivalent popula-tion of the two spin states for spin-1/2 nuclei. In addition, the sensitivity of theexperiment is decreased with pharmaceutical compounds, since they are com-posed primarily of carbon atoms where the 13C observable nuclei have a naturalabundance of only 1%. The long relaxation times of specific carbon nuclei alsopose a problem since quick, repetitive pulsing cannot occur. The technique ofcross-polarization provides a means of both signal enhancement and reductionof long relaxation times.

G. Cross-Polarization

The concept of cross-polarization as applied to solid-state NMR was implementedby Pines et al. (16). A basic description of the technique is the enhancement ofthe magnetization of the rare spin system by transfer of magnetization from theabundant spin system. Typically, the rare spin system is classified as 13C nucleiand the abundant system as 1H spins. This is especially the case for pharmaceuti-cal solids, and the remaining discussion of cross-polarization focuses on thesetwo spin systems only.

Figure 4 describes the cross-polarization (CP) pulse sequence and the be-havior of the 1H and 13C spin magnetizations during the pulse sequence in termsof the rotating frame of reference (19). Step 1 of the sequence involves rotationof the proton magnetization onto the y′ axis by application of a 90° pulse (rotat-ing-frame magnetic field B1H). Subsequent spin-locking occurs along y′ by anon-resonance pulse along y′ for a specific period of time, t. At this point, a high

Fig. 3 Solid-state 31P NMR spectra of fosinopril sodium acquired under single pulse,high-power proton decoupling and various conditions of magic-angle spinning: (A) static;(B) 2.5 kHz; (C) 4.0 kHz; (D) 5.0 kHz; (E) 6.0 kHz. The isotropic chemical shift is desig-nated by an asterisk. (From Ref. 11.)

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 12: Dk1719 ch10

Fig. 4 The top diagram represents the pulse sequence for the cross-polarization experi-ment, whereas the bottom diagram describes the behavior of the 1H and 13C spin magnetiza-tions during the sequence. (From Ref. 11.)

degree of proton polarization occurs along B1H, which will decay with a specifictime referred to as T1ρH. As the 1H spins are locked along y′, an on-resonancepulse, B1C, is applied to the 13C spins. The 13C spins are also ‘‘locked’’ along y′and decay with the time T1ρC, (step 2). By correctly choosing the magnitude ofthe spinlocking fields B1H and B1C, the Hartmann-Hahn (20) condition [Eq. (8)]will be satisfied and transfer of magnetization will occur from the 1H spin

γ1H B1H � γ13C B1C (8)

reservoir to the 13C spin reservoir. Once this equality is obtained by the correctspin-locking fields, the dilute 13C spins will take on the characteristics of the morefavorable 1H system. In the end, a maximum magnetization enhancement equalto the ratio of the magnetogyric ratios may be achieved (γH/γC). Once the carbonmagnetization has built up during this ‘‘contact period’’ (t), B1C is switched offand the carbon-free induction decay recorded (step 3). During this data acquisi-tion period, the proton field is maintained on for heteronuclear decoupling of the1H–13C dipolar interactions. Step 4 involves a standard delay period in which no

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 13: Dk1719 ch10

pulses occur and the two spin systems are allowed to relax back to their equilib-rium states.

The entire CP pulse sequence provides two advantages for the solid-stateNMR spectroscopist: significant enhancement of the rare spin magnetization, andreduction of the delay time between successive pulse sequences since the rarespin system takes on the relaxation character of the abundant spin system. Signalenhancement (magnetization enhancement) for less sensitive nuclei is immedi-ately apparent from the CP process. Less obvious is the fact that the rare spinsystem signal is not dependent on the recycle time in step 4 (regrowth of carbonmagnetization), but on the transfer process in step 2 and the relaxation behaviorof the proton spin system in step 4. Since the single spin lattice relaxation timeof the proton system is typically much shorter than the carbon system (1’s to10’s versus 100’s of seconds), the delay time is much shorter. This correspondsto a greater number of scans per unit time, yielding better S/N.

Throughout the cross-polarization pulse sequence, a number of competingrelaxation processes are occurring simultaneously. Recognition and understand-ing of these relaxation processes are critical in order to apply CP pulse sequencesfor quantitative solid-state NMR data acquisition or ascertaining molecular mo-tions occurring in the solid state.

H. Relaxation

Familiar to most chemists is the notion of spin-lattice relaxation (21). Labeledas T1, the spin-lattice relaxation time is defined as the amount of time for the netmagnetization (Mz) to return to its equilibrium state (M0) after a spin transitionis induced by a radiofrequency pulse [Eq. (9)]. The ‘‘lattice’’ term originatesfrom the idea that the spin system gives up energy to its surroundings as

∂Mz

∂ t

��1T1

(Mz � M0) (9)

it tries to reestablish spin equilibrium. The process of transferring spin energyto other modes of energy may be classified as relaxation mechanism. For T1 insolids, all of the following mechanisms may contribute: dipole–dipole (T1DD),spin-rotation (T1SR), quadrupolar (T1Q), scalar (T1SC), and chemical shift anisot-ropy (T1CSA). The simple inversion–recovery pulse sequence may be used to mea-sure T1 times for solid state samples (22). In the use of simple high-powereddecoupling, single-pulse sequences for quantitative NMR studies, the inversion–recovery pulse sequence may be used to determine T1’s of interest. A multipleof five times T1 may then be incorporated as the recycle time between successivepulses, assuring sufficient time for the magnetization to return to equilibrium. In

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 14: Dk1719 ch10

this way, the NMR signal observed is truly representative of the number of nucleiproducing it.

To understand the cross-polarization process, two other rate processes mustbe defined: (a) spin-lattice relaxation in the rotating frame (T1ρ) and (b) the cross-polarization relaxation time (TCH). The spin-lattice relaxation in the rotating framecharacterizes the decay of magnetization in a field B1, which is normally muchsmaller than the externally applied field B0. In steps 1 and 2 of the CP sequence(Fig. 4), the carbon and proton spin systems are locked by the application offields B1H and B1C and each system decays with its characteristic time. Figure 5represents a thermodynamic model for the relaxation processes during CP. Duringcontact between the two spin systems, magnetization is transferred at the rateTCH. Since competing relaxation processes are occurring, the following conditionsmust be met to obtain a spectrum by CP: T1C T1H T1ρH CP time TCH.It is apparent that for quantitative NMR studies of solids by CP, the individualrelaxation processes must be measured to assure that the signal is truly propor-tional to the amount of species present.

With an understanding of magnetic interactions in the solid state, inherentrelaxation processes, and experimental techniques to overcome these difficulties,

Fig. 5 Thermodynamic model of the cross-polarization sequence and representation ofthe competing relaxation processes. (From Ref. 11.)

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 15: Dk1719 ch10

Schaefer et al. acquired the first ‘‘liquid-like’’ NMR spectrum of a solid by CP/MAS techniques (23).

Before a review of recent applications of solid-state NMR to the pharma-ceutical sciences, a brief Experimental section is presented making reference toa number of important pulse sequences.

III. EXPERIMENTAL ASPECTS

Manufacturers of today’s modern NMR spectrometers normally offer a numberof different models of instruments that are capable of measuring solid-state NMRspectra. The basic components of the solid-state spectrometer are the same asthe solution-phase instrument: data system, pulse programmer, observe and de-coupler transmitters, magnetic system, and probes. In addition, high-power ampli-fiers are required for the two transmitters and a pneumatic spinning unit to achievethe necessary spin rates for MAS. Normally, the observe transmitter for 13C workrequires broadband amplification of approximately 400 W of power for a 5.87-T, 250-MHz instrument. The amplifier should have triggering capabilities so thatonly the radiofrequency (rf) pulse is amplified. This will minimize noise contribu-tions to the measured spectrum. So that the Hartmann-Hahn condition may beachieved, the decoupler amplifier must produce an rf signal at one-fourth thepower level of the observe channel for carbon work.

Although a series of standard referencing compounds has not been estab-lished for solid-state NMR as opposed to solution-phase studies (24), a numberof unofficial standard samples are currently being used (25). Table 2 lists a seriesof compounds, their intended use, and appropriate references further detailingtheir intended uses and limitations. Additionally, like solution-phase NMR, amyriad of solid-state NMR pulse sequences exist that selectively investigate thesample of interest. A number of these pulse sequences are described within thisExperimental section with further discussion outlined in the Applications sectionof this work.

The single pulse excitation/magic-angle spinning (SPE/MAS) pulse se-quence is a means of obtaining solid-state NMR spectra of very sensitive nucleisuch as phosphorus-31. The sequence is an adaptation of the solution-phase, gateddecoupling NMR experiment (32). In the SPE/MAS sequence, high-power protondecoupling occurs during the signal acquisition period. This decoupling periodprimarily removes the heteronuclear dipolar coupling that typically broadenssolid-state NMR spectra. This pulse sequence is normally used to acquire decou-pled spectra on heteronuclear atoms (signified as X in the pulse sequence) thatpossess high sensitivity (such as the 31P spectra presented in Fig. 3). MAS is usedthroughout the sequence to minimize broadening due to CSA.

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 16: Dk1719 ch10

Table 2 Reference Samples for Solid-State NMR

Compound Intended use Reference

Adamantane (C10H16) External referencing for 13C spec- (26)tra, optimizing Hartmann-Hahnmatch, linewidth measurement,sensitivity measurement

Hexamethylbenzene (C12H18) Optimizing Hartmann-Hahn match (26)Glycine (C2H5O2N) External referencing for 13C spec- (26)

tra, sensitivity measurement,magic angle setting for 13C exper-imentation

Ammonium phosphate External referencing for 31P spectra (26)(NH4H2PO4) Monobasic

Potassium bromide (KBr) Magic angle setting for 13C experi- (27)mentation

ZNPa, Zn[S2P(OC2H5)2]2 Magic angle setting for 31P experi- (28)mentation

DMPPOb, C8H11OP Magic angle setting for 31P experi- (29)mentation

Samarium acetate tetrahydrate 13C CP/MAS chemical shift ther- (30)[(CH3CO2)3SmC4H2O] mometer

TTAAc 15N CP/MAS chemical shift ther- (31)mometer

a ZNP: zinc(II) bis (O,O′-diethyldithiophosphate).b DMPPO: dimethylphenylphosphine oxide.c TTAA: 1,8-dihydro-5,7,12,14-tetramethyldibenso (b,i)-1,4,8,11-tetraazacyclotetradeca-4,6,11,13-

tetraene-15N4 (tetramethyldibenzotetraaza(14)annulene).Source: Adapted from Ref. 11.

As demonstrated in Fig. 3, even with high-speed MAS, spinning sidebandsdo occur. These sidebands may be confused with actual resonances in the NMRspectrum and be misinterpreted. Dixon et al. have developed a pulse sequencewhich suppresses spinning sidebands typically observed in NMR spectra acquiredwith high-field magnets operating at 5 T and above (18). This pulse sequencehas been labeled TOSS, total suppression of sidebands. After initial cross-polarization, Hahn spin echoes are generated with a series of pulses on the Xchannel. The timing of these pulses are programmed such that the sidebands areinverted, but the isotropic peak is unaffected. Upon addition of spectra with in-verted and normal sideband intensities, the sidebands cancel and the isotropicpeak remains.

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 17: Dk1719 ch10

Another pulse sequence that is helpful for the assignment of resonances ina qualitative CP/MAS NMR analysis is the delayed decoupling, or dipolar de-phasing experiment. This pulse sequence is simply the CP/MAS experiment witha variable delay time inserted between the spin-locking period and the start ofdecoupling/acquisition. Nonprotonated carbons are selectively observed by thispulse sequence (33), due to the inherent weaker 13C–1H dipolar interaction forcarbons with nonbonded protons. During the variable delay time (usually pro-grammed between 50 and 100 µs), directly protonated carbons dephase rapidlyand their signal disappears. Subsequently, a simplified spectrum is measured inwhich only quaternary carbons are observed.

Quantitative analysis is a very important component of pharmaceuticalanalysis. In a number of the publications describing the use of solid-state NMR,the majority of the work has dealt with qualitative studies with brief referencesto the possibility of quantitative analysis. Under proper data acquisition condi-tions, solid-state NMR is a quantitative technique that typically provides suffi-cient selectivity and sensitivity. An excellent guide to the utilization of magicangle spinning and cross-polarization techniques for quantitative solid-state NMRdata acquisition has been outlined by Harris (34).

As discussed in the Theory section, in order to acquire solid-state NMRspectra in which the signal intensities truly reflect the nuclei producing it, dataacquisition parameters such as recycle time, pulse widths, cross-polarization time,Hartmann-Hahn match, and decoupling power must be explicitly determined foreach chemical system. If these mechanisms, primarily relaxation, are not under-stood and improper acquisition parameters inserted into the pulse sequence, thequantitative nature of solid-state NMR can be compromised. In Harris’s article,the problems of solid-state NMR acquisition techniques as applied to quantitativemeasurements are addressed. Additionally, errors in the setting of the magicangle, Hartmann-Hahn match, and cross-polarization mixing time are discussedin relation to obtaining quantitative NMR results. To this end, a series of pulsesequences is presented here in order to determine the time factor of various relax-ation mechanisms and to program this information into the quantitative data ac-quisition pulse sequence.

From previous discussions it is known that competing rate processes occurin the cross-polarization NMR experiment. So that the measured NMR signaltruly represents the number of nuclei producing it, explicit measurements mustbe made for these various mechanisms, namely, TCP and T1ρH. For SPE experi-ments, T1 must be measured for the nucleus of interest.

During the Hartmann-Hahn condition, proton magnetization is being trans-ferred to the carbon spins at the rate of 1/TCH and to the lattice at the rate of1/T1ρH. In order for CP to occur, 1/TCH � 1/T1ρH (1/T1ρC is being ignored, sinceit is greater than 1/T1ρH). The 13C magnetization will achieve a maximum value

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 18: Dk1719 ch10

at a specific contact time, which also signifies optimum sensitivity for the CPexperiment. For any quantitative analysis studies, the optimal contact time shouldbe determined for the chemical system of interest. Normally, the contact time ismeasured by varying the contact time in the CP/MAS pulse sequence and measur-ing the subsequent signal intensity. By plotting the contact time versus the signalintensity on a logarithmic scale, the optimal contact time for each resonance maybe determined (35). Once the contact time is determined for the distinct resonanceobserved in the NMR spectrum, this value, or a compromised/average value,may be inserted into the quantitative CP/MAS pulse sequence.

In addition to measuring TCH for the chemical system in question, the protonT1ρ value must be determined since the repetition rate of a CP experiment isdependent on the recovery of the proton magnetization. Common conventionstates that a delay time between successive pulses of 1–5 � T1 must be used.One advantage to solids NMR work is that a common proton T1ρ value will bemeasured, since protons communicate through a spin-diffusion process. The mostcommon way to measure the proton T1ρ is to apply a proton 90° pulse, wait avariable delay time, and then apply the normal CP pulse sequence. By measuringthe 13C intensity versus the variable delay time, the proton T1ρ time may be ap-proximated (35).

Once the proton T1ρ and TCH values are determined for the chemical system,physical mixtures of the two entities can be generated (calibration samples). Sub-sequent acquisition of the solid-state NMR spectra under quantitative conditionsyields signal intensities representative of the amount of each material present.Measurement of the relative signal intensities may then lead to determination ofthe amount of one species present in an unknown mixture.

In some cases, CP is not necessary to obtain a suitable solid-state NMRspectrum. In these cases, the SPE/MAS sequence may be used and for quantita-tive analysis only the X-nucleus T1 time needs to be determined. The standardinversion–recovery experiment can be used to measure this value, keeping inmind that MAS and high-power decoupling is still necessary. As before, oncethe X-nucleus T1 time is determined, 1–5 � T1 may be inserted as the delay timebetween successive pulses for quantitative data acquisition.

IV. PHARMACEUTICAL APPLICATIONS

Solid-state NMR has and continues to be used in a myriad of pharmaceuticalapplications, but unfortunately, all of the different applications cannot be dis-cussed in this chapter. Instead, applications in drug development, as opposed todrug discovery, are presented herein. Aboul-Enein has outlined some of the gen-eral uses of solid-state NMR for pharmaceutical research (36). Phosphorus-31

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 19: Dk1719 ch10

cross-polarization/magic-angle spinning (CP/MAS) NMR has been used to studythe dehydration of disodium clodronate (37). A fast rise in temperature revealsthat disodium clodronate loses lattice water and only one 31P resonance is mea-sured, whereas a slow increase in temperature converts the crystalline form toan anhydrous form which displays two nonequivalent phosphorus atoms (reso-nances).

Solid-state NMR has been used extensively for characterizing the structuresof N-desmethylnefopam HCl (38), patellin (39), erythromycin A dihydrate (40),and the amorphous nature of ursodeoxycholic acid (41). The conformations of3′-amino-3′-deoxythymidine (42), gramicidin A (43), and amiodarone HCl (44)have been confirmed by solid-state NMR.

Due to the specificity of solid-state NMR, it is an ideal technique to studyinclusion complexes, drug–excipient interactions, or the effect of moisture onthe drug substance or formulation. Studies on inclusion complexes includegliclazide-β-cyclodextrin (45) and hydrocortisone butyrate (46), whereas Makriy-annis utilized NMR for investigating drug–membrane interactions (47). Otherinteractions studied by solid-state NMR include polyethylene glycol with griseo-fulvin (48) and the moisture-induced interaction with trospectomycin sulfate,which affects the 3′-gem-diol ↔ 3′-keto equilibrium in either the drug substanceor freeze-dried formulation (49).

A. Polymorphism

The most widely used application of solid-state NMR in drug development is inthe area of polymorphism (50,51). Pharmaceutical solids can exist in a numberof solid forms, each having different properties of pharmaceutical importance,including stability and bioavailability; the number and properties of these formsare largely unpredictable and vary considerably from case to case. Pharmaceuticalsolids can be divided into crystalline and amorphous solids based on the presenceof crystalline structure. This classification is usually based on X-ray powder dif-fraction and/or microscopic examination. Crystalline solids can then be furtherclassified into: crystalline polymorphs, forms having the same chemical composi-tion but different crystal structures and therefore different densities, meltingpoints, solubilities, and many other important properties; and solvates, forms con-taining solvent molecules within the crystal structure, giving rise to unique differ-ences in solubility, response to atmospheric moisture, loss of solvent, etc. Some-times a drug substance may be a desolvated solvate, which is formed whensolvent is removed from a specific solvate while the crystal structure is essentiallyretained—again, many important properties are unique to such a form.

Different physical forms of a drug substance can display radically differentsolubilities, which directly affects the dissolution and bioavailability characteris-

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 20: Dk1719 ch10

tics of the compound. In addition, the chemical stability of one form, as comparedto the other, may vary. The physical stability is also crucial. During various pro-cessing steps (grinding, mixing, tablet pressing, etc.), the physical form of thedrug substance may be compromised, subsequently leading to dissolution prob-lems. For these reasons, the full characterization of polymorphic systems is criti-cal to numerous entities (preformulation/physical pharmacy group, chemical pro-cess development, regulatory/patent office, and analytical laboratories) withindrug development companies.

The use of solid-state NMR for the investigation of polymorphism is easilyunderstood based on the following model. If a compound exists in two true poly-morphic forms, labeled as A and B, each crystalline form is crystallographicallydifferent. This means, for instance, that a carbon nucleus in form A may be situ-ated in a slightly different molecular environment as compared to the same carbonnucleus in form B. Although the connectivity of the carbon nucleus is the samein each form, the local environment may be different. Since the local environmentmay be different, this potentially leads to a different chemical-shift interactionfor each carbon, and ultimately, a different isotropic chemical shift for the samecarbon atom in the two different polymorphic forms. If one is able to obtain purematerial for the two forms, analysis and spectral assignment of the solid-stateNMR spectra of the two forms can lead to the origin of the crystallographicdifferences in the two polymorphs. Solid-state NMR is thus an important tool inthe multidisciplinary approach to polymorphism.

Solid-state NMR analysis of different solid forms can provide crucial infor-mation which often resolves conflicting results between other methods such asinfrared, X-ray powder diffraction, and thermal methods (52). In addition, a dic-tate of formulation technology is that the physical form of the drug substance,after being defined and verified, should not change once the product has beenmanufactured. Solid-state NMR provides a powerful method for comparing thephysical form of the drug substance after pharmaceutical processing or manufac-turing. Furthermore, solid-state NMR provides a method for the analysis of mix-tures of crystal forms in both the pure drug substance and in dosage forms(52).

There are a number of significant advantages to the use of solid-state NMRfor the study of polymorphism. As compared to diffuse reflectance IR, Raman,and X-ray powder diffraction techniques, solid-state NMR is a bulk techniquein which the intensity of the measured signal is not a function of the particle sizeof the material. In addition, NMR is an absolute technique under proper dataacquisition procedures, meaning that the intensity of the signal is directly propor-tional to the number of nuclei producing it, thus permitting quantitative analysis.In some polymorphic investigations, a single crystal of sufficient quality maynot be available for X-ray crystallography (structure determination). By proper

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 21: Dk1719 ch10

assignment of the NMR spectrum, though, the origin of the polymorphism maybe inferred by differences in the resonance position for identical nuclei in eachpolymorphic form.

B. Drug Substance Qualitative Analysis

The majority of applications of solid-state NMR used in the investigation ofpharmaceutical polymorphs are performed in conjunction with other analyticaltechniques. Byrn et al. have reported differences in the solid-state NMR spec-tra for different polymorphic forms of benoxaprofen, nabilione, and pseudo-polymorphic forms of cefazolin (53). Although single-crystal X-ray crystallogra-phy was initially used to study the polymorphs, the solid-state 13C CP/MAS NMRspectrum of each form was distinctly different. These studies focused primarilyon the bulk drug material, although a granulation of benoxaprofen was studied.It was concluded that solid-state NMR could be used to differentiate the formpresent in the granulation even in the presence of excipients. In further studies,the crystalline forms of prednisolone tert-butylacetate (54), cefaclor dihydrate(55), and glyburide (56) were studied. The five crystal forms of prednisolonetert-butylacetate were again determined by single-crystal X-ray crystallography,and solid-state NMR was used to determine the effect of crystal packing on the13C chemical shifts of the different steroid forms. Although conformationalchanges were observed in the ester side chain by X-ray crystallography, no majordifferences were noted in the NMR spectra, indicating that the environment re-mains relatively unchanged. Significant chemical shift differences were noted forcarbonyl atoms involved in hydrogen bonding. This theme is consistent in theNMR study of cefaclor dihydrate. Again, the effects of hydrogen bonding werediscernible by solid-state NMR. The study of glyburide was principally con-cerned with the structural conformation of the molecule in the solution and solidstate. The solution conformation was determined by 1H and 13C NMR, and in thesolid by single-crystal X-ray crystallography, IR, and solid-state 13C NMR. Thesolid-state NMR results suggested that this method would be useful for compar-ing solid- and solution-state conformations of molecules.

In 1989, Etter published an excellent paper on the complementary use ofsolid-state NMR and X-ray crystallography to study the structure of pharmaceuti-cal solids (57). Crystallographic effects such as polymorphism, multiple mole-cules per asymmetric unit cell, disorder, intra- and intermolecular H-bonding,tautomerism, and solvation were all investigated by solid-state NMR.

In a series of publications by Harris and Fletton, solid-state NMR has beenused to investigate the structure of polymorphs. The majority of studies involveX-ray and IR techniques and also address the possibility of quantitative measure-ments of polymorphic mixtures. The three pseudo-polymorphic forms of testos-

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 22: Dk1719 ch10

terone were examined by IR and 13C CP/MAS NMR (58). The two forms ofmolecular spectroscopy were able to differentiate the forms, but NMR had theability to investigate nonequivalent molecules in a given unit cell. A series ofdoublet resonances was noted for a series of different carbon atoms. This impliedthat the specific carbon atom within the molecule may resonate at two differentfrequencies, depending on the crystallographic site of the molecule. In additionto the hydrogen-bonding explanation of the crystallographic splittings, the useof NMR to determine quantitatively the amount of each pseudo-polymorph pres-ent in a mixture was addressed. In the study of androstanolone (59), high-quality13C NMR spectra were obtained by CP/MAS techniques and allowed for charac-terization of the anhydrous and monohydrate forms. Again, crystallographic split-tings were noted for the two forms and were related to hydrogen bonding. Anidentical approach to study pharmaceutical polymorphic structure was used inthe investigation of the two polymorphs of 4′-methyl-2′-nitroacetanilide (60). Anadditional study of cortisone acetate (61,62) by solid-state NMR revealed differ-ences in the NMR spectra for the six crystalline forms. Further multidisciplinaryapproaches to the physical characterization of pharmaceutical compounds aredetailed in separate studies on cyclopenthiazide (63), the excipient lactose (64),fursemide (65), losartan (66), fosinopril sodium (67), and leukotriene antagonistsMK-679, MK-571 (68), L-660,711 (69), captopril (70), diphenhydramine HCl(71), mofebutazone (72), phenylbutazone (72), oxyphenbutazone (72), cimetidine(73), and the iron chelator 1,2-dimethyl-3-hydroxy-4-pyridone (74). In each case,solid-state NMR was used in conjunction with other techniques such as DSC,IR, X-ray diffraction, microscopy, and solubility/dissolution studies to character-ize the polymorphic systems fully.

In further studies of polymorphism by solid-state NMR, conversion fromone solid-state form to another by UV irradiation (75) and variable temperaturetechniques (76) are outlined. In the first study, NMR was employed to followthe chemical transformation within the organic crystals of p-formyl-trans-cinnamic acid (p-FCA). A photoreactive β-phase may be crystallized from etha-nol, whereas a photostable γ-phase is produced from acetone. After irradiationof the β-phase with UV radiation, and subsequent acquisition of the solid-state13C NMR spectrum, the photoproduct was easily identified by NMR. The secondconversion study (76) investigated the four forms of p-amino-benzenesulphona-mide sulphanilamide (α, β, γ, and δ). The first three forms were fully investi-gated by solid-state 13C NMR and X-ray crystallography techniques. Subsequentvariable-temperature studies monitored the interconversion of the α- and β-formsto the γ-form. Coalescence of some NMR signals in the γ-form also suggestedthat phenyl ring motion occurred within the crystal. Conclusions from the studyindicated that solid-state NMR had the ability to differentiate pharmaceuticalpolymorphs, determine asymmetry in the unit cell, and investigate molecular mo-tion within the solid state.

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 23: Dk1719 ch10

C. Drug Substance Quantitative Analysis

In a number of the publications describing the use of solid-state NMR for poly-morphic characterization, the majority of the work has dealt with qualitative stud-ies, with brief references to the possibility of quantitative analysis. Quantitativesolid-state 13C CP/MAS NMR has been used to determine the relative amountsof carbamazepine anhydrate and carbamazepine dihydrate in mixtures (77). The13C-NMR spectra for the two forms did not appear different, although sufficientS/N for the spectrum of the anhydrous form required long accumulation times.This was determined to be due to the slow proton relaxation rate for this form.Utilizing the fact that different proton spin-lattice relaxation times exist for thetwo different pseudopolymorphic forms, a quantitative method was developed.The dihydrate form displayed a relatively short relaxation time, permitting in-terpulse delay times of only 10 s to obtain full-intensity spectra of the dihydrateform while displaying no signal due to the anhydrous form. By utilizing an inter-nal standard (glycine) and the differences in the relaxation rate of the two forms,the peak area of the dihydrate could be measured and related through a calibrationcurve to the amount of anhydrous and dihydrate content in mixtures of carbama-zepine.

D. Drug Product Analysis

A series of papers has been published on the solid-state NMR spectra of a numberof analgesic drugs. Jagannathan recorded the solid-state 13C NMR spectrum ofacetaminophen in bulk and dosage forms (78). From the solution-phase NMRspectrum, assignments of the solid-state NMR resonances could be made in addi-tion to explanations for the doublet structure of some resonances (dipolar cou-pling). Spectra of the dosage product from two sources indicated identical drugsubstance, but different levels of excipients. The topic of drug–excipient interac-tions was addressed by solid-state 13C NMR in the investigation of different com-mercially available aspirin samples (79,80). In each commercial aspirin product,the only difference in the measured NMR spectrum was due to variations in theexcipients, indicating that there were no interactions between the drug and theexcipients under dry blending conditions. After lyophilization of two of the prod-ucts, one aspirin sample did show a different NMR spectrum, indicating possibleinteraction during lyophilization or conversion to a different solid-state form dur-ing processing.

Saindon et al. used solid-state NMR to show that different brands of pred-nisolone tablets contained different polymorphs (81). For the listed drugs, theNMR analysis could discern the active component except for the low-dose enala-pril maleate tablets. Excipients obscure some drug resonances which appear atthe same chemical shift value, but overall, have little effect on resonances at

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 24: Dk1719 ch10

other chemical shifts. Similar results were obtained for tablets of another closelyrelated HMG-CoA reductase inhibitor, simvastatin.

E. Mobility

Molecular mobility can be studied by solid-state NMR and X-ray crystallography.Solid-state NMR offers several approaches to studying the molecular mobilityof solids. These include: (a) the study of processes which result in peak coales-cence of solid-state NMR resonances using variable temperature solid-stateNMR, (b) determination of the T1 relaxation of individual carbon atoms usingvariable temperature solid-state NMR, (c) use of interrupted decoupling to detectmethylene and possible methine groups with unusual mobility, and (d) compari-son of solid-state MAS spectra measured with and without cross-polarization.

The rate of relaxation of the carbon spins toward equilibrium is character-ized by the spin-lattice relaxation time, T1 (82). In this process, the excess energyfrom the spin system is given to the surroundings or lattice. Interactions of ran-domly fluctuating magnetic fields at the Lamor frequency of the nucleus stimu-lates these transitions. Such fields arise from motions of other nuclear magneticmoments such as protons. Spin-lattice relaxation is thus most efficient when thesemotions are near the Lamor frequency. Studies of T1 can give valuable informa-tion about fast motions in the megahertz range, such as methyl group rotations.Variable-temperature solid-state NMR spectroscopy has been shown to be a pow-erful technique for the study of molecular motion in cortisone (83), cholesterol(84), proteins (85), polymers (86–90), and amino acids (91).

In addition, solid-state NMR has been used to study protein hydration andstability (92) and the activation energies of spinning methyl groups (93). Theactivation energy for the spinning of methyl groups in triethylphosphine oxidewas reported to be 7.9 kJ/mol by T1 studies and 8.69 kJ/mol by T1ρ studies (88).Relaxation studies of peri-substituted naphthalenes were used to determine thebarrier to methyl rotation when substituents at the peri position were EH, ECH3,ECl, and EBr (94). The observed barriers were EH, 9.7 kJ/mol; ECH3, 13.5kJ/mol; ECl, 15.2 kJ/mol; and EBr, 18.4 kJ/mol. The barriers follow the se-quence of the van der Waals radii and indicate that if a large group abuts themethyl group, it can restrict methyl rotation. Rotation barriers for methyl groupsin several amino acids have also been reported (93).

F. Chirality

The issue of chirality continues to be an important issue in drug development.With the advent of solid-state NMR, enantiomeric purity of a sample may bemeasured directly. A number of reports in the literature demonstrate the abilityof 13C or 31P solid-state NMR to differentiate between optically pure material and

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 25: Dk1719 ch10

racemic species (95). In one report, tartaric acid was studied by 13C CP/MASNMR (96). For each of the optically pure (2R, 3R), racemic, and meso-tartaricacid material studied, two molecules exist per asymmetric unit cell. Since thecarbonyls and α-carbon atoms in a single molecule are not symmetry related,one would expect two resonances for each carbon (crystallographic splitting). The13C NMR spectra showed two resonances for each carbon, but also demonstrated adistinct isotropic chemical shift value for each resonance upon comparison toeach other. This data demonstrated the potential for solid-state NMR to differenti-ate between racemic and enantiomeric crystallites.

A series of racemic and optically pure organophosphorus samples werestudied by solid-state 31P NMR (97). Analogous to the tartaric acid work, the twooptical isomers (� and �) showed the same 31P NMR spectrum that were distinctfrom the spectrum of the racemic material. In addition, the issue of quantitativeanalysis of optical purity was addressed. The conformational analysis of dl-, l-,and d-methionine was studied by solid-state 13C NMR (98). Based on the NMRstudies, each crystalline form of dl-methionine consists of a single conformer,which agrees with X-ray diffraction data.

G. Future Direction

In recent years, two-dimensional solid-state NMR spectroscopy (99,100) has beenapplied to pharmaceutical systems. The overall goal of these studies is to obtainstructural information. Such studies are particularly important and useful whenthe crystal structure cannot be determined.

Raftery’s laboratory has used two-dimensional (2D) solid-state NMR tostudy the red, orange, and yellow conformational polymorphs of ROY (5-methyl-2-(2-nitrophenyl)amino]-3-thiophenecarbonitrile (101). The separation of side-bands by order experiment (2DTOSS) separates the isotropic and anisotropicchemical shifts over two dimensions, allowing analysis of the chemical-shift an-isotropy patterns using magic-angle spinning in the 1500-Hz range (102). Thechemical shift anisotropy of the C-3 carbon atom (the atom attached to the CN)could be easily distinguished using this pulse sequence and was used to probedifferences in chemical environment between the different color polymorphs.Raftery and co-workers showed that two-dimensional solid-state NMR can easilydistinguish differences between conformational polymorphs and is a powerfulmethod for analyzing such systems. Raftery’s group is now combining informa-tion from the 2DTOSS experiments with ab-initio or density functional methodsto attempt to obtain quantitative structural information. This combined approachalong with other 2-D solid-state NMR experiments, promises to provide structuralinformation on polymorphs whose crystal structure cannot be determined.

Munson and co-workers have used 2D, high-speed 13C CP/MAS NMR tostudy uniformly 13C-labeled aspartame (103). Aspartame exists in three hydrates

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 26: Dk1719 ch10

(two hemihydrates and a dihemihydrate). Two of these hydrates apparently con-tain three molecules per asymmetric unit, giving up to three resolvable resonancesfor each carbon atom. This complex crystal packing prevented assignments ofthe resonances using techniques based on the number of attached protons or Jcouplings. Typical MAS experiments at spinning rates of 7 kHz and proton de-coupling powers of 63 kHz gave broadened spectra due to broadening by dipolardecoupling. However, increasing the spinning rate to 28 kHz and the decouplingto 263 kHz gave narrow line spectra that allowed the assignment of the crystallo-graphically inequivalent sites. For one of the forms, peak assignments could bemade for all three molecules in the asymmetric unit of the crystal.

As with the 2DTOSS studies, Munson and co-workers showed that solid-state NMR can provide important information on the structures of polymorphsand solvates whose crystal structure is not known. As we seek more detailedstructural information of drugs in order to understand more about solid-state be-havior, these 2D techniques will no doubt become more and more important.Additional 2D techniques have been utilized to study DNA complexes (104)and perform distance measurements (105), specifically using the rotational-echodouble-resonance (REDOR) sequence (106,107).

Regulatory documentation is now making specific reference to solid-stateNMR. A flow chart approach to the physical characterization of pharmaceuticalsolids was first published in 1995 (108). In the flow chart approach to determiningthe number of polymorphic forms, spectroscopy, and specifically solid-stateNMR, is outlined as a recommended technique. The U.S. Food and Drug Admin-istration (FDA) has also recognized the need for solid-state NMR characterizationof drug substance or product. In the most recent International Committee of Har-monization (ICH) guideline for the setting of specifications for drug substanceand product, spectroscopy (IR, Raman, and solid-state NMR) is referred to indecision tree four: investigating the need to set acceptance criteria for polymor-phism in drug substances and drug products (109). Worldwide regulatory authori-ties now recognize the importance of solid-state NMR and how the technique isintimately tied to the drug development process.

V. SUMMARY

The multinuclear technique of solid-state NMR has been applied to various as-pects of drug development with tremendous success. Although the technique ide-ally lends itself to the structure determination of drug compounds in the solidstate, it is anticipated that in the future, solid-state NMR will be used routinely formethod development and problem-solving activities in the analytical/materialsscience/physical pharmacy area of the pharmaceutical sciences. During the pastfew years, an increasing number of publications have emerged in which solid-

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 27: Dk1719 ch10

state NMR has become an invaluable technique. With the continuing develop-ment of solid-state NMR pulse sequences and hardware improvements (increasedsensitivity), solid-state NMR will provide a wealth of information for the charac-terization of pharmaceutical solids.

REFERENCES

1. PA Mirau. A strategy for NMR structure determination. J Magn Reson 96:480–490, 1992.

2. AH Beckelt. Chirality and its importance in drug development. Biochem Soc Trans19:443–446, 1991.

3. O Kaplan, JS Cohen. Magnetic resonance as a non-invasive tool to study metabo-lism in pharmacological research. Trends Pharmacol Sci 11:398–399, 1990.

4. FMH Jeffrey, A Rajagopal, CR Malloy, AD Sherry. 13C-NMR: A simple yet com-prehensive method for analysis of intermediary metabolism. Trends Biochem Sci16:5–10, 1991.

5. V Stoven, JY Lallemand, D Abergel, S Bovaziz, MA Delsuc, A Ekondzi, E Guittet,S Laplante, R LeGoas, T Malliavin, A Mikov, C Reisdorf, M Rogin, C van Heijen-oort, Y Yang. Insight into protein nuclear magnetic resonance research. Biochimie72:531–535, 1990.

6. JPM van Duynhoven, PJM Folkers, CWJM Prinse, BJM Harmsen, RNH Konings,CW Hilbers. Assignment of the 1H NMR spectrum and secondary structure elucida-tion of the single-stranded DNA binding protein encoded by the filamentous bacte-riophage IKe. Biochemistry 31:1254–1262, 1992.

7. GM Clore, AM Gronenborn. Determination of structures of larger proteins in solu-tion by three- and four-dimensional heteronuclear magnetic resonance spectros-copy. In: GM Clore, AM Gronenborn, eds. NMR of Proteins. Boca Raton, FL:CRC Press, 1993, pp. 1–32.

8. JK Baker, CW Myers. One-dimensional and two-dimensional 1H- and 13C-nuclearmagnetic resonance (NMR) analysis of vitamin E raw materials or analytical refer-ence standards. Pharm Res 8:763–770, 1991.

9. A Burger, R Ramberger. On the polymorphism of pharmaceuticals and other molec-ular crystals. I. Theory of thermodynamic rules. Mikrochim Acta [Wien] II:259–271, 1979. A Burger, R Ramberger. On the polymorphism of pharmaceuticals andother molecular crystals. II. Applicability of thermodynamic rules. Mikrochim Acta[Wien] II:273–316, 1979.

10. J Haleblian, W McCrone. Pharmaceutical applications of polymorphism. J PharmSci 58:911–929, 1969.

11. DE Bugay. Solid state nuclear magnetic resonance spectroscopy—theory and phar-maceutical applications. Pharm Res 10:317–327, 1993.

12. Karplus MJ. Vicinal proton coupling in nuclear magnetic resonance. J Am ChemSoc 85:2870–2871, 1963.

13. TC Farrar, ED Becker. Pulse and Fourier Transform NMR. New York: AcademicPress, 1971.

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 28: Dk1719 ch10

14. ER Andrew, A Bradbury, RG Eades. Removal of dipolar broadening of nuclear mag-netic resonance spectra of solids by specimen rotation. Nature 183:1802–1803, 1959.

15. ER Andrew. The narrowing of NMR spectra of solids by high-speed specimenrotation and the resolution of chemical shift and spin multiplet structures for solids.Prog Nuclear Magn Reson Spectrosc 8:1–39, 1972.

16. A Pines, MG Gibby, JS Waugh. Proton-enhanced nuclear induction spectroscopy.A method for high resolution NMR of dilute spins in solids. J Chem Phys 56:1776–1777, 1972. A Pines, MG Gibby, JS Waugh. Proton-enhanced NMR of dilute spinsin solids. J Chem Phys 59:569–590, 1973.

17. A Salvetti. Newer ACE inhibitors, a look at the future. Drugs 40:800–828, 1990.18. WT Dixon, J Schaefer, MD Sefcik, EO Stejskal, RA McKay. Total suppression of

sidebands in CPMAS carbon-13 NMR. J Magn Reson 49:341–345, 1982.19. CA Fyfe. Solid State NMR for Chemists. Guelph, Ontario, Canada: CFC Press,

1983.20. SR Hartmann, EL Hahn. Nuclear double resonance in the rotating frame. Phys Rev

128:2042–2053, 1962.21. RJ Abraham, J Fisher, P Loftus. Introduction to NMR Spectroscopy. New York:

Wiley, 1988:84–86.22. TC Farrar, ED Becker. Pulse and Fourier Transform NMR. New York: Academic

Press, 1971:20–22.23. J Schaefer, EO Stejskal, R Buchdahl. High-resolution carbon-13 nuclear magnetic

resonance study of some solid, glassy polymers. Macromolecules 8:291–296, 1975.24. P Granger. Multinuclear NMR referencing. Appl Spectrosc 42:1–3, 1988.25. WL Earl, DL VanderHart. Measurement of 13C chemical shifts in solids. J Magn

Reson 48:35–54, 1982.26. Bruker Instruments. The CP/MAS Accessory Product Description Manual. Biller-

ica, MA: Bruker Instruments, 1987.27. JS Frye, GE Maciel. Setting the magic angle using a quadrupolar nuclide. J Magn

Reson 48:125–131, 1982.28. A Kubo, CA McDowell. Setting of the magic angle for 31P MAS NMR using

zinc(II) bis(O, O′-diethyldithiophosphate). J Magn Reson 92:409–410, 1991.29. RC Crosby, JF Haw. Dimethylphenylphosphine oxide, a compound for setting the

magic angle in 31P CP/MAS NMR. J Magn Reson 82:367–368, 1989.30. GC Campbell, RC Crosby, JF Haw. 13C chemical shifts which obey the curie law

in CP/MAS NMR spectra. The first CP/MAS NMR chemical-shift thermometer.J Magn Reson 69:191–195, 1986.

31. B Wehrle, F Aguilar-Parrilla, H-H Limbach. A novel 15N chemical-shift NMR ther-mometer for magic angle spinning experiments. J Magn Reson 87:584–591, 1990.

32. RJ Abraham, J Fisher, P Loftus. Introduction to NMR Spectroscopy, New York:Wiley, pp. 113–116, 1988.

33. SJ Opella, MH Frey. Selection of nonprotonated carbon resonances in solid-statenuclear magnetic resonance. J Am Chem Soc 101:5854–5856, 1979.

34. RK Harris. Quantitative aspects of high-resolution solid-state nuclear magnetic res-onance spectroscopy. Analyst 10:649–655, 1985.

35. D Michel, F Engelke. Cross-Polarization, Relaxation Times and Spin-Diffusion inRotating Solids. In: P Diehl, E Fluck, H Gunther, R Kosfeld, J Seelig, eds. NMR

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 29: Dk1719 ch10

Basic Principles and Progress, Vol. 32. New York: Springer-Verlag, 1994, pp. 69–125.

36. Y Aboul-Enein. Applications of solid-state nuclear magnetic resonance spectros-copy to pharmaceutical research. Spectroscopy 5:32–33, 1990.

37. JT Timonen, E Pohjala, H Nikander, TT Pakkanen. A 31P CP-MAS NMR studyon dehydration of disodium clodronate tetrahydrate. Pharm Res 15:110–115, 1998.

38. R Glaser, S Geresh, J Blumenfeld, D Donnell, N Sugisaka, M Drouin, A Michel.Solution and solid-state structures of N-desmethylnefopam hydrochloride, a metab-olite of the analgesic drug. J Pharm Sci 82:276–281, 1993.

39. TM Zabriskie, MP Foster, TJ Stout, J Clardy, CM Ireland. Studies on the solution-and solid-state structure of patellin 2. J Am Chem Soc 112:8080–8084, 1990.

40. GA Stephenson, JG Stowell, PH Toma, RR Pfeiffer, SR Byrn. Solid-state investiga-tions of erythromycin A dihydrate: Structure, NMR spectroscopy, and hygroscopic-ity. J Pharm Sci 86:1239–1244, 1997.

41. E Yonemochi, Y Ueno, T Ohmae, T Oguchi, S-I Nakajima, K Yamamoto. Evalua-tion of amorphous ursodeoxycholic acid by thermal methods. Pharm Res 14:798–803, 1997.

42. T Kovacs, L Parkanyi, I Pelczer, F Cervantes-Lee, KH Pannell, PF Torrence. Solid-state and solution conformation of 3′ amino-3′-deoxythymidine precursor to a non-competitive inhibitor of HIV-1 reverse transcriptase. J Med Chem 34:2595–2600,1991.

43. RR Ketchem, W Hu, A Cross. High-resolution conformation of gramicidin A in alipid bilayer by solid-state NMR. Science 261:1457–1460, 1993.

44. RC Rao, F Lefebvre, G Commenges, C Castet, L Miguel, G Maire, M Gachon.Conformation of amiodarone HCl: Solution and solid state 13C NMR study. PharmRes 11:1088–1092, 1994.

45. JR Moyano, MJ Arias-Blanco, JM Gines, F Giordano. Solid-state characterizationand dissolution characteristics of gliclazide-β-cyclodextrin inclusion complexes. IntJ Pharm 148:211–217, 1997.

46. IK Chun, DS Yun. Inclusion complexation of hydrocortisone butyrate with cyclo-dextrins and dimethyl-β-cyclodextrin in aqueous solution and in solid state. Int JPharm 96:91–103, 1993.

47. A Makriyannis, T Mavromoustakos. Studies on drug-membrane interactions usingsolid state NMR, small angle X-ray diffraction, and differential scanning calorime-try. Epitheor Klin Farmakol Farmakokinet, Int Ed 3:95–114, 1989.

48. M Alden, J Tegenfeldt, ES Saers. Structures formed by interactions in solid disper-sions of the system polyethylene glycol griseofulvin with charged and non-chargedsurfactants added. Int J Pharm 94:31–38, 1993.

49. MD Likar, RJ Taylor, PE Fagerness, Y Hiyama, RH Robins. The 3′-keto-diolequilibrium of trospectomycin sulfate bulk drug and freeze-dried formulation:solid-state carbon-13 cross-polarization magic angle spinning (CP/MAS) and high-resolution carbon-13 nuclear magnetic resonance (NMR) spectroscopy studies.Pharm Res 10:75–79, 1993.

50. DE Bugay. Magnetic Resonance Spectrometry. In: HG Brittain, ed. Physical Char-acterization of Pharmaceutical Solids. New York: Marcel Dekker, pp. 93–125,1995.

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 30: Dk1719 ch10

51. SR Byrn, B Tobias, D Kessler, J Frye, P Sutton, P Saindon, J Kozlowski. Relation-ship between solid state NMR spectra and crystal structures of polymorphs andsolvates of drugs. Trans Am Crystallogr Assoc 24:41–54, 1989.

52. DE Bugay. Solid state nuclear magnetic resonance spectroscopy—theory and phar-maceutical applications. Pharm Res 10:317–327, 1993.

53. SR Byrn, G Gray, RR Pfeiffer, J Frye. Analysis of solid-state carbon-13 NMRspectra of polymorphs (benoxaprofen and nabilone) and pseudopolymorphs (cefaz-olin). J Pharm Sci 74:565–568, 1985.

54. SR Byrn, PA Sutton, B Tobias, J Frye, P Main. Crystal structure, solid-state NMRspectra, and oxygen reactivity of five crystal forms of prednisolone tert-butylace-tate. J Am Chem Soc 110:1609–1614, 1988.

55. H Martinez, SR Byrn, RR Pfeiffer. Solid-state chemistry and crystal structure ofcefaclor dihydrate. Pharm Res 7:147–153, 1990.

56. SR Byrn, AT McKenzie, MMA Hassan, AA Al-Badr. Conformation of glyburidein the solid state and in solution. J Pharm Sci 75:596–600, 1986.

57. MC Etter, GM Vojta. The use of solid-state NMR and X-ray crystallography ascomplementary tools for studying molecular recognition. J Mol Graphics 7:3–11,1989.

58. RA Fletton, RK Harris, AM Kenwright, RW Lancaster, KJ Packer, N Sheppard. Acomparative spectroscopic investigation of three pseudopolymorphs of testosteroneusing solid-state i.r. and high-resolution solid-state NMR. Spectrochim Acta 43A:1111–1120, 1987.

59. RK Harris, BJ Say, RR Yeung, RA Fletton, RW Lancaster. Cross-polarization/magic-angle spinning NMR studies of polymorphism: androstanolone. SpectrochimActa 45A:465–469, 1989.

60. RA Fletton, RW Lancaster, RK Harris, AM Kenwright, KJ Packer, DN Waters, AYeadon. A comparative spectroscopic investigation of two polymorphs of 4′-methyl-2′-nitroacetanilide using solid-state infrared and high-resolution solid-statenuclear magnetic resonance spectroscopy. J Chem Soc Perkin Trans 2, 11:1705–1709, 1986.

61. RK Harris, AM Kenwright, BJ Say, RR Yeung, RA Fletton, RW Lancaster, GLHardgrove Jr. Cross-polarization/magic-angle spinning NMR studies of polymor-phism: Cortisone acetate. Spectrochim Acta 46A:927–935, 1990.

62. EA Christopher, RK Harris, RA Fletton. Assignments of solid-state 13C resonancesfor polymorphs of cortisone acetate using shielding tensor components. Solid StateNuclear Magn Reson 1:93–101, 1992.

63. JJ Gerber, JG van derWatt, AP Lotter. Physical characterisation of solid forms ofcyclopenthiazide. Int J Pharm 73:137–145, 1991.

64. HG Brittain, SJ Bogdanowich, DE Bugay, J DeVincentis, G Lewen, AW Newman.Physical characterization of pharmaceutical solids. Pharm Res 8:963–973, 1991.

65. C Doherty, P York. Frusemide crystal forms; solid state and physicochemical analy-ses. Int J Pharm 47:141–155, 1988.

66. K Raghavan, A Dwivedi, GC Campbell Jr, E Johnston, D Levorse, J McCauley,M Hussain. A spectroscopic investigation of losartan polymorphs. Pharm Res 10:900–904, 1993.

67. HG Brittain, KR Morris, DE Bugay, AB Thakur, ATM Serajuddin. Solid-state

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 31: Dk1719 ch10

NMR and IR for the analysis of pharmaceutical solids: polymorphs of fosinoprilsodium. J Pharm Biomed Anal 11:1063–1069, 1994.

68. RG Ball, MW Baum. A spectroscopic and crystallographic investigation of thestructure and hydrogen bonding properties of the chiral leukotriene antagonist MK-679 as compared to its racemate MK-571. J Org Chem 57:801–803, 1992.

69. EB Vadas, P Toma, G Zografi. Solid-state phase transitions initiated by water vaporsorption of crystalline L-660, 711, a leukotriene D4 receptor antagonist. Pharm Res8: 148–155, 1991.

70. HY Aboul-Enein, RA Dommisse, HO Desseyn. Solid state NMR of captopril andits zinc comples. Spectrosc Lett 23:491–495, 1990.

71. R Glaser, K Maartmann-Moe. X-ray crystallography studies and CP-MAS 13CNMR spectroscopy on the solid-state stereochemistry of diphenhydramine hy-drochloride, an antihistaminic drug. J Chem Soc Perkin Trans 2:1205–1210,1990.

72. M Stoltz, DW Oliver, PL Wessels, AA Chalmers. High-resolution solid statecarbon-13 nuclear magnetic resonance spectra of mofebutazone, phenylbutazone,and oxyphenbutazone in relation to X-ray crystallographic data. J Pharm Sci 80:357–362, 1991.

73. DA Middleton, CS Duff, F Berst, DG Reid. Cross-polarization magic-angle spin-ning 13C NMR characterization of the stable solid state forms of cimetidine. J PharmSci 86:1400–1402, 1997.

74. H-K Chan, S Venkataram, DJW Grant, Y-E Rahman. Solid state properties of anoral iron chelator, 1,2-dimethyl-3-hydroxy-4-pyridone and its acetic acid solvate.I: Physicochemical characterization intrinsic dissolution rate and solution thermo-dynamics. J Pharm Sci 80:677–685, 1991.

75. KDM Harris, JM Thomas. Probing polymorphism and reactivity in the organicsolid state using 13C NMR spectroscopy: Studies of p-formyl-trans-cinnamic acid.J Solid State Chem 93:197–205, 1991.

76. L Frydman, AC Olivieri, LE Diaz, B Frydman, A Schmidt, S Vega. A 13C solid-state NMR study of the structure and the dynamics of the polymorphs of sulphani-lamide. Mol Phys 70:563–579, 1990.

77. R Suryanarayanan, TS Wiedmann. Quantitation of the relative amounts of anhy-drous carbamazepine (C15H12N2O) and carbamazepine dihydrate (C15H12N2O⋅2H2O)in a mixture by solid-state nuclear magnetic resonance (NMR). Pharm Res 7:184–187, 1990.

78. NR Jagannathan. High-resolution solid-state carbon-13 nuclear magnetic resonancestudy of acetaminophen: A common analgesic drug. Curr Sci 56:827–830, 1987.

79. C Chang, LE Dıaz, F Morin, DM Grant. Solid state 13C NMR study of drugs: Aspi-rin. Magn Res Chem 24:768–771, 1986.

80. LE Dıaz, L Frydman, AC Olivieri, B Frydman. Solid state NMR of drugs: Solubleaspirin. Anal Lett 20:1657–1666, 1987.

81. PJ Saindon, NS Cauchon, PA Sutton, C-J Chang, GE Peck, SR Byrn. Solid-statenuclear magnetic resonance (NMR) spectra of pharmaceutical dosage forms. PharmRes 10:197–203, 1993.

82. E Fukushima, SBW Roeder. Experimental Pulse NMR: A Nuts and Bolts Ap-proach. Reading, MA: Addison-Wesley, 1981.

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 32: Dk1719 ch10

83. ER Andrew, M Kempka. Proton NMR study of molecular motion in solid cortisone.Solid State Nuclear Magn Reson 2:261–264, 1993.

84. ER Andrew, B Peplinska. NMR study of solid cholesterol. Mol Phys 70:505–512,1990.

85. S Yoshioka, Y Aso, S Kojima. Determination of molecular mobility of lyophilizedbovine serum albumin and γ-globulin by solid-state 1H NMR and relation toaggregation-susceptibility. Pharm Res 13:926–930, 1996.

86. MD Poliks, J Schaefer. Microscopic dynamics in chloral polycarbonate by cross-polarization magic-angle spinning carbon-13 NMR. Macromolecules 23:2682–2686, 1990.

87. J Schaefer, MD Sefcik, EO Stejskal, RA McKay, WT Dixon, RE Cais. Molecularmotion in glassy polystyrenes. Macromolecules 17:1107–1118, 1984.

88. J Schaefer, MD Sefcik, EO Stejskal, RA McKay. Carbon-13 T1ρ experiments onsolid polymers having tightly spin-coupled protons. Macromolecules 17:1118–1124, 1984.

89. AN Garroway, WM Ritchey, WB Moniz. Some molecular motions in epoxy poly-mers: A carbon-13 solid-state NMR study. Macromolecules 15:1051–1063, 1982.

90. J Schaefer, EO Stejskal, R Buchdahl. Magic-angle carbon-13 NMR analysis ofmotion in solid glassy polymers. Macromolecules 1982:384–405, 1977.

91. A Naito, S Ganapathy, K Akasaka, CA McDowell. Spin-relaxation of carbon-13solid amino acids using the CP-MAS technique. J Magn Reson 54:226–235, 1983.

92. F Separovic, YH Lam, X Ke, H-K Chan. A solid-state NMR study of protein hydra-tion and stability. Pharm Res 15:1816–1821, 1998.

93. ER Andrew, WS Hinshaw, MG Hutchins, ROI Sjoblom. Proton magnetic relaxationand molecular motion in polycrystalline amino acids. I. Aspartic acid, cystine, gly-cine, histidine, serine, tryptophan, and tyrosine. Mol Phys 31:1479–1488, 1976.

94. F Imashiro, K Takegoshi, S Okazawa, J Furukawa, T Terao, A Saika, A Kawamori.NMR study of molecular motion in perisubstituted naphthalenes. J Chem Phys 78:1104–1111, 1983.

95. ZJ Li, MT Zell, EJ Munson, DJW Grant. Characterization of racemic species ofchiral drugs using thermal analysis, thermodynamic calculation, and structural stud-ies. J Pharm Sci 88:337–346, 1999.

96. HDW Hill, AP Zens, J Jacobus. Solid-state NMR spectroscopy. Distinction of dia-stereomers and determination of optical purity. J Am Chem Soc 101:7090–7091,1979.

97. KV Anderson, H Bildsøe, HJ Jakobsen. Determination of enantiomeric purity fromsolid-state 31P NMR of organophosphorus compounds. Magn Res Chem 28:S47–S51, 1990.

98. LE Dıaz, F Morin, CL Mayne, DM Grant, C-J Chang. Conformational analysis ofDL-, L-, and D-methionine by solid-state 13C NMR spectroscopy. Magn ReasonChem 24:167–170, 1986.

99. DP Burum, A Bielecki. An improved experiment for heteronuclear-correlation 2DNMR in solids. J Magn Reson 94:645–652, 1991.

100. A Bielecki, DP Burum, DM Rice, FE Karasz. Solid-state two-dimensional13C-1H correlation (HETCOR) NMR spectrum of amorphous poly (2,6-dimethyl-p-phenylene oxide) (PPO). Macromolecules 24:4820–4822, 1991.

Copyright 2002 by Marcel Dekker. All Rights Reserved.

Page 33: Dk1719 ch10

101. J Smith, E MacNamara, D Raftery, T Borchardt, S Byrn. Application of two-dimensional 13C solid-state NMR to the study of conformational polymorphism. JAm Chem Soc 120:11710–11713, 1998.

102. SF deLacroix, JJ Titman, A Hagemeyer, HW Spiess. Increased resolution in MASNMR spectra by two-dimensional separation of sidebands by order. J Magn Res97:435–443, 1992.

103. MT Zell, BE Padden, DJW Grant, MC Chapeau, I Prakash, EJ Munson. Two-dimensional high-speed CP/MAS NMR spectroscopy of polymorphs. 1. Uniformly13C-labeled aspartame. J Am Chem Soc 121:1372–1378, 1999.

104. C-L Juang, RA Santos, P Tang, W-J Chien, M-H Wann, R Bernstein, J Hong, GSHarbison. Structure and dynamics of DNA and drug DNA complexes by one andtwo-dimensional solid-state NMR. J Biomol Struct Dyn 8:18–22, 1991.

105. TP Jarvie, JS Bader, GT Went. Method and apparatus for distance measurementswith solid-state NMR and usefulness for drug design. U.S. Patent Number 9743627, 1997.

106. DD Beusen, LM McDowell, U Slomczynska, J Schaefer. Solid-state nuclear mag-netic resonance analysis of the conformation of an inhibitor bound to thermolysin.J Med Chem 38:2742–2747, 1995.

107. TP Jarvie, GT Went, KT Mueller. Simultaneous multiple distance measurementsin peptides via solid-state NMR. J Am Chem Soc 118:5330–5331, 1996.

108. S Byrn, R Pfeiffer, M Ganey, C Hoiberg, G Poochikian. Pharmaceuticals solids:A strategic approach to regulatory considerations. Pharm Res 12: 945–954, 1995.

109. Q6A Specifications: Test Procedures and Acceptance Criteria for New Drug Sub-stances and New Drug Products: Chemical Substances. Fed Reg 62, No. 22762890–62910, 1997.

Copyright 2002 by Marcel Dekker. All Rights Reserved.