Field Cycling in NMR Relaxation Spectroscopy: Applications ...

24
Field Cycling in NMR Relaxation Spectroscopy: Applications in Biological, Chemical and Polymer Physics Rainer Kimmich Sektion Kernresonanzspektroskopie UniversitSt Ulm D-7900 Ulm, West Germany I. INTRODUCTION 195 II. PRINCIPAL ASPECTS OF NUCLEAR MAGNETIC RELAXATION 195 A. General Remarks 195 B. Relaxation Mechanisms 197 C. Correlation and Intensity Functions 198 D. Heterogeneously Relaxing Systems 200 III. FIELD CYCLING EXPERIMENTS 201 A. Why field cycling? 201 B. Longitudinal Relaxation 203 C. Transverse Relaxation 204 D. Experimental Requirements and Limits 204 IV. APPLICATIONS 207 A. Biological Systems 207 B. Cells and Tissues 211 C. Liquid Crystals . < 211 D. Synthetic Oligomers and Polymers 212 E. Non-macromolecular Systems with Electron- paramagnetic Centers 215 V. FINAL REMARKS 216 References 216 I. INTRODUCTION Nuclear magnetic relaxation spectroscopy has the po- tential to be a very versatile method for the investigation of molecular dynamics in combination with structural details. This fact has been well known since the early days of NMR (1). The successful exploitation of many possible applications, however, required technical progress which has only occurred in the last few years. This article deals with the special technique of field cycling in which the NMR signal is detected in a reasona- bly high field after a relaxation period during which the magnetic field strength was switched to a lower or higher value. (Typical field cycles are described in section III.) Thus, the relaxation can be studied in variable magnetic fields without varying the frequency of the spectrometer. In other words, the cyclic variation of the external mag- netic flux density B o increases the available range by several orders of magnitude. Advances in high power electronics and the technology of cryomagnets allow the measurement of relaxation rates, often including in- dividual intensity functions, in the 10- 5 to 1.0 T field range, corresponding to a proton frequency range of 10 3 to 10 8 Hz with any kind of proton-containing systems. For X-nuclei the frequency range is shifted according to the different magnetogyric ratio. Thus, the necessary tool is now available to distinguish different types of molecular dynamics. On the other hand, theoretical models of molecular dynamics have been developed, which indicate sig- nificant differences between relaxation spectra. Several authors have established catalogues of correlation func- tions appropriate for the discussion of many problems. Thus, both the apparatus and the theoretical back- ground are available to promise solutions of problems in molecular dynamics. The following two sections review experimental and theoretical fundamentals and describe the limits of the method at the present state of the art and developments expected in the future. In section IV, the applications performed to date by several groups are classified and discussed. We shall restrict ourselves to applications relevant to biological, chemical, and polymer physics. II. PRINCIPAL ASPECTS OF NUCLEAR MAGNETIC RELAXATION A. General Remarks The theory of the nuclear magnetic relaxation times T-, and T 2 has been treated in different ways with different ranges of validity. Reviews are given in several text- books (2-4). In this section, we will restrict ourselves to those parts of the relaxation theories predominantly relevant to the applications to be subsequently outlined. With regard to the applicability of perturbation theory, one distinguishes between the cases of "weak" and Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal. Vol. 1,No. 4 195

Transcript of Field Cycling in NMR Relaxation Spectroscopy: Applications ...

Page 1: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

Field Cycling in NMR Relaxation Spectroscopy:Applications in Biological, Chemical

and Polymer Physics

Rainer KimmichSektion Kernresonanzspektroskopie

UniversitSt UlmD-7900 Ulm, West Germany

I. INTRODUCTION 195II. PRINCIPAL ASPECTS OF NUCLEAR MAGNETIC RELAXATION 195

A. General Remarks 195B. Relaxation Mechanisms 197C. Correlation and Intensity Functions 198D. Heterogeneously Relaxing Systems 200

III. FIELD CYCLING EXPERIMENTS 201A. Why field cycling? 201B. Longitudinal Relaxation 203C. Transverse Relaxation 204D. Experimental Requirements and Limits 204

IV. APPLICATIONS 207

A. Biological Systems 207B. Cells and Tissues 211C. Liquid Crystals .< 211D. Synthetic Oligomers and Polymers 212E. Non-macromolecular Systems with Electron-

paramagnetic Centers 215V. FINAL REMARKS 216References 216

I. INTRODUCTION

Nuclear magnetic relaxation spectroscopy has the po-tential to be a very versatile method for the investigationof molecular dynamics in combination with structuraldetails. This fact has been well known since the earlydays of NMR (1). The successful exploitation of manypossible applications, however, required technicalprogress which has only occurred in the last few years.

This article deals with the special technique of fieldcycling in which the NMR signal is detected in a reasona-bly high field after a relaxation period during which themagnetic field strength was switched to a lower or highervalue. (Typical field cycles are described in section III.)Thus, the relaxation can be studied in variable magneticfields without varying the frequency of the spectrometer.In other words, the cyclic variation of the external mag-netic flux density Bo increases the available range byseveral orders of magnitude. Advances in high powerelectronics and the technology of cryomagnets allow the

measurement of relaxation rates, often including in-dividual intensity functions, in the 10-5 to 1.0 T fieldrange, corresponding to a proton frequency range of 103

to 108 Hz with any kind of proton-containing systems.For X-nuclei the frequency range is shifted according tothe different magnetogyric ratio. Thus, the necessarytool is now available to distinguish different types ofmolecular dynamics.

On the other hand, theoretical models of moleculardynamics have been developed, which indicate sig-nificant differences between relaxation spectra. Severalauthors have established catalogues of correlation func-tions appropriate for the discussion of many problems.Thus, both the apparatus and the theoretical back-ground are available to promise solutions of problems inmolecular dynamics.

The following two sections review experimental andtheoretical fundamentals and describe the limits of themethod at the present state of the art and developmentsexpected in the future. In section IV, the applicationsperformed to date by several groups are classified anddiscussed. We shall restrict ourselves to applicationsrelevant to biological, chemical, and polymer physics.

II. PRINCIPAL ASPECTS OF NUCLEARMAGNETIC RELAXATION

A. General Remarks

The theory of the nuclear magnetic relaxation times T-,and T2 has been treated in different ways with differentranges of validity. Reviews are given in several text-books (2-4). In this section, we will restrict ourselves tothose parts of the relaxation theories predominantlyrelevant to the applications to be subsequently outlined.

With regard to the applicability of perturbation theory,one distinguishes between the cases of "weak" and

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

Vol. 1,No. 4 195

Page 2: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

"strong" collisions (5). The papers reviewed in this articlemainly concern the weak-collision case where theZeeman field-splitting is large compared with the jn-fluence of the perturbing field.

There are two types of weak-collision theories. Thefirst implies that the spin system reaches a spin tempera-ture (i.e., a Boltzmann distribution of the level popula-tions combined with a completely random distribution ofthe Larmor precession phases) in a time short comparedwith 7Y This condition is usually fulfilled in solids, forwhich the Gorter-Hebel-Slichter equation holds in thehigh-temperature limit (reference 3, p. 143).

1 _ 1T-, 2

Wmn (Em - Enf

£ E2

where Wmn is the probability per second that a transitiontakes place from the state m to the state n characterizedby the spin energies Em and En, respectively. The disad-vantage of this type of theory is that it is restricted tosystems which rapidly reach a spin temperature, but ithas the advantage that the quantum mechanical fea-tures of the lattice can be taken into account. This playsan important role with relaxation mechanisms due to thecoupling of nuclei to conduction electrons (ref. 2, p. 355)or to spin-phonon coupling (ref. 2, p. 401).

In other cases, a theoretical perturbation treatmentwith no explicit consideration of the lattice quantummechanics is preferable. Then, a semi-classical treat-ment is performed taking the lattice globally into ac-count. Several modifications of this type of theory havebeen outlined in the literature and lead to the same re-sults. A straightforward derivation of the most importantstandard case (dipolar coupling in two-spin systems) hasbeen given (6) to which most of the following applicationsare related.

For nuclei of spin 1/2, we have to deal with the patternof transitions between the eigenstates of the longitudinaland transverse components of the spin operators asgiven in Figure 1. The relaxation rates are then given forlike spins, i.e., for equal magnetogyric ratios (y, = ys), inthe high-temperature limit by (6)

^ w2) (2a)

and in the motional-narrowing case by1

-zr= 2(Ui + U2)' 2

For unlike spins (7, =£ 7S)

• ~ w0 + 2M' 1

and (again in the motional-narrowing case)

—= u0 + 2u1 + u2>2

(2b)

w2 (2c)

(2d)

(A) (B)

Figure 1. Transition probabilities between eigenstates of thespin operators: (A) longitudinal components, (B) transversecomponents. (6)

In the derivation of equations 2c and 2d, it was as-sumed that the nonresonant spin species stay always inequilibrium as a consequence of an additional interac-tion of the S-spins with the lattice. This condition is oftenfulfilled for unpaired electrons (zero-field-splitting) andfor quadrupole nuclei (coupling to the electric-fieldgradient of the lattice).

The transition probabilities in equations 2a-2d can becalculated with the aid of first-order perturbation theory(6) or its equivalent modifications (ref. 2, p. 264 and ref.3, p. 137). The essential result is that the probabilitiescan be expressed by the intensity functions of the fluc-tuating interaction. In the standard case mentionedabove, the probability w (Figure 1), for instance, is givenby (6)

Ah2

where a = - -^ - ^ h2 y, ys.

(3)

li (cc) is the intensity function to be taken at the reso-nance frequence w and defined by the Wiener-Khintchine relation

where G^ (T) is the autocorrelation function of the part ofthe dipolar interaction which induces the flip of the singlespin

Gi(r) = F,(0) F](T) (5)

R, <f>, 6 are the time-dependent spherical coordinates ofthe spin-spin vector. The bar denotes the ensembleaverage over all pairs. The expressions for the othertransition probabilities are summarized elsewhere (6).Thus, the correlation functions play a central role in all

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

196 Bulletin of Magnetic Resonance

Page 3: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

weak-collision cases. Instead of the correlation functionin equation 5, it is more convenient to introduce thereduced, or normalized, correlation function

G(r) =F,(0) F{{r) -

(6)

with the mean square deviation

<T;= \F/ 2 _

so that G(0) = 1 and G(°°) = 0.The Fourier transform of G(r) will be designated by

l(w). For isotropic systems defined by the condition F, =0, equation 6 can be simplified further. Equating the en-semble average with the time average for a single spinpair, it becomes clear that the term "isotropic system"means that during the relaxation times each interactionsituation has occurred with the same probability.

Frequently one has to deal with more than two inter-action partners. It is then a good approximation to cal-culate the effective relaxation rates (7"12~

1)eff as a sumover the two-spin rates r,^"1 of all spins interacting witha reference nucleus, even if the spin states are perturbedin a correlated manner (2,7,8)

narrowing effect is incomplete, and this may be the caseeven for very rapid but anisotropic motions, the Bloch-decay of the transverse magnetization can be describedby the Anderson-Weiss formula (9,10)

M,rans( t) = MUans(0) exp { - M/ j (t - r) gu (T) df] (8): O

where M2r is the second moment of the rigid lattice and

gjj) is the autocorrelation function of the fluctuatingpart of the resonance frequency.

When the externally applied magnetic field Bo ap-proaches zero, the strong collision case becomesrelevant. The spin-lattice relaxation time is then (5)

.2(1 -P) £BO (9)

— )= E

7"i,2eff pairs1 (7)

This two-spin approximation is generally assumed in theliterature.

Transverse relaxation in highly viscous or solidlikesystems implies some complication. If the motional-

where T* is the mean time between two "collisions" (i.e.,atomic rearrangements), p is a function of the dipolarfields changed with a collision (p = 1, if no change oc-curs; p = 0 if there is any correlation to the fields beforethe collision is lost). BL is the local field caused by themagnetic dipoles.

B. Relaxation Mechanisms

The standard case of dipolar interaction is the relaxa-tion mechanism most frequently assumed in NMR ap-plications. There is, however, a series of quite differentinteractions leading to nuclear magnetic relaxationwhich should always be kept in mind. Table 1 gives a

Table 1. Nuclear Magnetic Relaxation Mechanisms

Spin of theRelaxing Nucleus

/ > y2

/> 1

Coupledby

dipolarinter-action

scalar in-teraction(Fermicon-tact andindirectcoupling)

chemicalshifttensor

spin ro-tationtensor

Coulombinterac-tion

to

equal d ipo le^

unequal d ipole^

dipole/I2ofaquadrupolenucleus

local magneticfield

angular momen-tum of themolecule

electric fieldgradient

Examples

1H**Mn2+

1H<-14N

13C**e0>67

gases

14N in asym-metric molecules

ModulationSource

fluctuations of?12

fluctuations ofT12. M2

fluctuations offi2. M2

molecular re-orientation

atomic collisions

molecular reorien-tation (latticevibrations)

References

2,6

2,11,12

13,14

2,15,16

2,17

2,18

Note: Quantities \x^ and M2are * n e dipole moments of the interacting particles; r12 is the distance vector.

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

Vol. 1,No. 4 197

Page 4: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

brief survey of the different mechanisms and the refer-ences in which they are described and quantitativelyexplained.

C. Correlation and Intensity Functions

The most important aim of field-cycling relaxationspectroscopy is to obtain information about intensityand correlation functions of microscopic fluctuations. Inorder to show the potential differences in the types offluctuations occurring in molecular dynamics, we willreview some simple model cases distinguishable by themethod. The high and low frequency limits are given in

Table 2. The models are suitable throughout for dipolarinteraction. Some of the more generalized models canalso be applied to the other relaxation mechanisms.

A frequently assumed but not necessarily justifiedstandard case is that of an exponential correlationfunction

G(r) = e x p ( - | T | / T C ) (10a)

1 +(10b)

where TC is the correlation time. There are two mainsituations for which these functions hold. The first is theisotropic and continuous rotational diffusion of a dipole

Table 2. Limiting Behavior of Several Fluctuation Models*

Type of fluctuation

Rotational diffusion(a) isotropic(b) anisotropic

Markov jump models fordiscrete interactionstates and environ-ments (c)

Translational diffusion(a) continuous(b) stepwise

modulating intermolecularinteractions

Reorientation by(a) continuous(b) stepwise

diffusion of stabledefects with restoringcapability

(c)in 1 dimension,unlimited

(d) in 3 dimensions,unlimited

(e) in 1 dimension,limited

(f) on a crystallattice

Reorientation by continuous,unlimited diffusion of stabledefects without restoringcapability

(a) in 1 dimension(b) in 3 dimensions

Reorientation by curvilinearchain diffusion

Low frequency/high temperature

WT; « 1

-1

7", -v E ax T;i

i

1/2 -1/2(C) / , -v O) To

(d.e.or^To"1

(a,b)T, -VTO"1

1/2 -1/2' 1 ~ W To

High frequency/low temperature

C0Ti» 1

i

_ 3/2 1/2

(a) r, -v a) T0

(b) T-i ~ a;2 T0

3/2 1/2(a) Tf ~ o) TO

3/2 1/2

(a) T, -v o) TO

(b) T-i ~ 0)2 T 0

3/2 1/2/ 1 ~ 0) To

References

(a) 1,2(b)19(c) 20,21

22

(a-e)23,24

(f) 25,26

24,27

28

*Time parameters T-, and TO indicate the time scales of molecular reorientations and displacements. The same symbols areused for all models for simplicity. Exact definitions vary with the model and can be found in the references. Quantities a, areweighting factors.

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal

198 Bulletin of Magnetic Resonance

Page 5: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

pair with a fixed distance (29 and ref. 2, p. 298). Thesecond case appears with Poisson-like jumps between adiscrete set of perturbation states of any source, char-acterized by a single jump rate (3,20,21,30-32). While it ishard to verify the first case experimentally becausemolecules are either not spherical or not rigid (implyinginternal motions), the second situation occurs, for in-stance, in solid benzene (31), with NH3 groups (33), andfor crystal water in gypsum (32). All examples showedexcellent agreement with the theory. Nevertheless, theapplicability of this simple correlation function remainsan exception.

The anisotropic, continuous, and rotational diffusionof ellipsoidal molecules bearing the interacting dipolesleads to a linear combination of several exponentials(19). Also compare with (34). The same holds for internalrotational diffusion about an axis fixed on a tumblingmolecule (35-37), for the exchange in two-phasesystems (38,39), and for Markovian jumps between adiscrete set of perturbation states of any source withdifferent jump rates (21). The latter situation can gener-ally be treated as a Markovian chain, the mathematics ofwhich can be found (40). In this-context, the influence ofenvironmental fluctuation (Figure 2) is of special interest(41). In cases where the analytical treatment fails, anumerical calculation can be used even for rather com-plicated models (23).

Figure 2. Two-state model showing the effect of fluctuationbetween different environments on the kinetics and degree ofmolecular reorlentation. (21)

Intermolecular interactions between dipoles can bemodulated by translational diffusion. If this contributionto nuclear magnetic relaxation is dominant, a quitedifferent type of correlation function occurs (22,42-44,ref. 2, p. 300). In the limit of continuous translational dif-fusion and of a continuous distribution of spins, we have

^j =J^) i /2 f^ (11a)

du+(o>2To

2/l/4)(11b)

with T0 = d2/2D. D is the translational diffusion coef-ficient, n is the number of spins per unit volume, d is theminimum distance of the interacting nuclei, and J3/2 is a

Bessel function. A modified theory for liquid crystals hasbeen given (45). The application of these formulae is re-stricted to cases where the influence of intramolecularinteraction is weak. An example would be moleculesbearing a single spin. An experimental distinctionbetween intra- and intermolecular interactions is possi-ble with the use of isotopically diluted systems.

In solid materials, especially if they are amorphous,the reorientation of molecules or of parts of moleculescan be coupled to the arrival of structural defects. Exam-ples are vacancies or those defects discussed in thepolymer section. Generally this type of fluctuation iscalled "defect diffusion." A non-Markovian behavior isthe consequence because then the probability for thereorientation of a molecular group at a definite timedepends on the positions of defects at a previous time.

Defect diffusion in crystalline materials has beentreated as a random walk on a special type of lattice(22,25,26). For the continuous diffusion of defects inone- and three-dimensional disordered systems, anearest-neighbor approximation has been suggested byGlarum (46) and by Hunt and Powles (47). This approx-imation is inconsistent with the assumption that there isno mutual hindrance of the defects. Perturbation by thenon-nearest neighbor defects can be excluded only inthe case of mutual hindrance, where the defects can notpermeate each other. In fact, different results have beenobtained by taking into account all defects and not onlynearest neighbors (24,27). Assuming that the arrival of adefect leaves an interaction state completely uncorrelat-ed with the initial state, gives in the case of unlimitedone-dimensional diffusion (24,27)

(12a)

[ l - S2(a)]}(12b).,3/2

1 - C2(a)]

with a = (4a)Tp)-1 and re = ir £0 (16 D)-\ where £o is themean nearest defect distance, D is the diffusion coef-ficient, and C2 and S2 are the Fresnel integrals. In threedimensions, an exponential correlation function againoccurs (27).

A different situation arises when the defects have re-storing capability. This means that after the passage ofthe defect the original interaction state is restored in-stead of an uncorrelated state as in the previous model.A molecular group then suffers a series of uncorrelated"pulses" due to the passing defects. Therefore, the rateof the pulses influences only the absolute value of therelaxation rate but not the shape of the intensity func-tion, in clear contrast to the "total correlation loss"models.

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

Vol. 1,No. 4 199

Page 6: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

For unlimited one-dimensional diffusion of non-interacting defects with restoring capability we obtain(24)

JJL1TTb

/(«>) =

(13a)

(13b)

- e x p ( - sin

where rb = bz/2D. The quantity b is the width of a defectand D is the diffusion coefficient. The symbol erf desig-nates the error function. The treatment has been ex-tended to three dimensions (24).

Mutual hindrance of the defects alters the situationespecially in the one-dimensional case where defectscannot "overtake" each other. One has then to considerthe fluctuations of the whole ensemble of defects. Thiscan be done by a computer simulation (48,49). An ana-lytic approximation can be achieved by treating the twoneighboring defects as reflecting barriers limiting themean square displacements (Figure 3). This has beenperformed for continuous (50) and for stepwise (23)diffusion.

The exact expressions (23,50) are lengthy. We wanttherefore to present only the limiting expressions for theintensity function derived from a 6th order approxima-tion (51). Two characteristic times play a role: the diffu-sion time rb fora mean square displacement b2 (b is thelength of a defect) and the diffusion time ra for a meansquare displacement d2 (d is the distance between thereflecting barriers). With these quantities we have

for wTd < 1

for 0)Tb <3C 1 <K U>Td

M = \/TbTd

1

(14a)

1-7= (14b)

O,3/2T-( / I £ - 1 )V T/,

Another application of the limited diffusion model will bedescribed in the model membrane section.

The defect diffusion models discussed so far assumethat the defect has an infinite lifetime, more precisely, alifetime long compared with the relevant diffusion times.At the opposite limit, where defects can be frequentlycreated and annihilated at any position in the system,the appearance of a perturbation by a defect at a

, reflecting barriers

bi"i

Figure 3. Schematic representation of defect diffusionbetween reflecting barriers at x = 0 and x = d. Reference nu-cleus is at x = r. Relaxation formulae must be averaged over allpossible values of r. Defect has width b. (50)

molecular group is then no longer diffusion controlled.The consequence is a Poisson-like fluctuation with anexponential correlation function (see section 3.1 of ref.21).

Molecular dynamics in polymer melts and concentrat-ed solutions are strongly influenced by a wormlike chaindiffusion process. This mechanism is described by (28)

I TG(T) = exp(^J - e r f (15a)

(15b)

with Tj> = C2/20. D is the curvilinear chain diffusion coef-ficient, £ the orientation correlation length of the diffu-sion path. This model is relevant for all systems wherethe reorientation of molecular groups is coupled to a dis-placement along a randomly curved diffusion path.

Oscillator models can be relevant in solids withfrozen-in molecular reorientation or diffusion. Stochas-tic excitations of vibrations by random pulses may thendescribe nuclear magnetic relaxation (13,30, chap. X ofref. 51). It is clear that the diverse oscillator models haveto be strictly distinguished from spin-phonon couplingprocesses (52).

This section indicated the broad variety of correlationand intensity functions which can be studied by the fieldcycling method in NMR relaxation spectroscopy. Thereis, however, an important condition for this type of study,namely the homogeneity of the relaxation in the system.This question will be discussed in the following section.

D. Heterogeneously Relaxing Systems

Samples with structural heterogeneity do notnecessarily relax in an inhomogeneous manner. Thequestion is whether the structural heterogeneity is sta-tionary or fluctuating. In the latter case, we have to takeinto account the time scale of relaxation itself. Spinsexperiencing the same type of fluctuation during the life-time of their states, relax equivalently. This means thatthere is either no structural heterogeneity or the

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

200 Bulletin of Magnetic Resonance

Page 7: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

structural inhomogeneity is fluctuating with a mean life-time essentially shorter than the lifetime of the spinstates (i.e., the relaxation times), so that all spins "see"the same type of fluctuation.

Spins in different chemical groups of stable com-pounds are permanently in different structural environ-ments, and therefore relax heterogeneously with differ-ent correlation fuctions. This behavior can easily be de-tected under high-resolution conditions (16,53). Spins inrapidly varying structural environments, implying, for in-stance, weak bonds which are frequently broken up,relax homogeneously.

Homogeneous relaxation means that a single correla-tion function is relevant, while heterogeneous relaxationrequires a discrete or continuous distribution of correla-tion functions. Several types of such distributions havebeen reviewed (54).

A problem may be that distributions of correlationfunctions do not always reveal themselves by a distribu-tion of relaxation times. Magnetization transfer by spindiffusion in solid-like substances (2,55-58) and cross-relaxation (59,60) can lead to average exponential relax-ation curves. Distributions of correlation functions cannevertheless be revealed by isotopic dilution of theresonant nuclei, so that the averaging mechanisms areprevented. Rapid material exchange, on the other hand,should be considered as a special case of environmentalfluctuation leading to a single correlation function(21,38,39,61).

It is clear that the occurrence of distributions of cor-relation functions, whether or not combined with a dis-tribution of relaxation times, causes a severe complica-tion of the analysis of relaxation data, apart from the ap-pearance of additional effects (62). The relatively longtime scale of nuclear magnetic relaxation times, how-ever, often allows the exclusion of such distributions inchemically homogeneous systems, even in cases whereno special experimental test is possible. In connectionwith the applications discussed below we will occasion-ally return to this question.

III. FIELD CYCLING EXPERIMENTSA. Why Field Cycling?

The conventional method to determine nuclear mag-netic relaxation times uses appropriate radio frequency(RF) pulse sequences and stationary magnetic fields(16). This method fails at low magnetic field strengthsbecause it becomes more and more difficult to detectNMR signals in decreasing magnetic fields due to theweaker nuclear induction, the lower Curie-magnetization, and the narrower bandwidth of the RFpulses at low frequencies. While it might be possible to

compensate for the poor sensitivity by larger samplevolumes and by signal accumulation, one cannotproduce high-power RF pulses short enough for the de-tection of solid state signals at frequencies below sever-al MHz.

The problem can partially be overcome by performingrelaxation experiments in the rotating frame (16,63-65).These methods, however, merely cover a frequencyrange below about 100 kHz, leaving a gap of two ordersof magnitude for solids between rotating-frame and con-ventional laboratory-frame experiments. Moreover, thetheory of the relaxation in the rotating frame is, thoughsimilar, not identical to that of the laboratory frame,leading to some inconvenience in the simultaneous in-terpretation of both experiments.

A more appropriate solution to the problem of mea-suring relaxation times in the whole range down to thekHz region is the field-cycling method. Typical fieldcycles are plotted in Figures 4 and 5. The general princi-ple is to combine the high sensitivity and bandwidth inhigh magnetic fields, where the sample is polarized andthe signal detected, with the relaxation effect in variablerelaxation fields.

Moreover, the increase in the available frequency(field) range is accompanied by two important advan-tages in the general measuring technique of relaxationtimes. First, the method is very rapid and allows automa-tization because, independent of the relaxation field, thedetection frequency is always the same, so that no timeconsuming adjustments of the RF apparatus are neces-sary for a varying relaxation field. Second, this techniqueis insensitive to systematic errors in 7^ caused by in-homogeneous RF fields (chap. 3 of ref. 16). The pertur-bation of the sample occurs by a rapid jump of the exter-nal field Bo and not by an RF pulse. As Bo is veryhomogeneous compared with RF fields, a homogeneousperturbation of the sample is guaranteed. In contrast tothe perturbation, the detection of free induction decaysby inhomogeneous RF pulses is uncritical and causes atmost a loss in signal intensity but no systematic error.

Before outlining the practical aspect of the technique,let us first classify the general aims of the possible ap-plications into three main groups.

1. Relaxation DispersionSince molecular dynamics are characterized by cer-

tain correlation functions leading to a variety of intensityfunctions, identification of a special type of fluctuationrequires a measurement of the intensity function over aswide a range as possible. This can be done by relaxationspectroscopy in two ways. One can shift the intensityfunction by changing the temperature to vary the fluc-tuation time constants while observing the relaxationrates at a certain frequency given by the spacing of the

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

Vol. 1, No. 4 201

Page 8: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

Figure 4. Typical field cycle (idealized) for the determination of 7"v The z-component of the magnetization is increased to the equili-brium value MXP in the polarization field Bof>. In the relaxation field Bo

r, the magnetization relaxes with T^B,/) toward Mxr. Detec-

tion occurs in the detection field 6od(here chosen equal to fio") with the aid of a 90° RF pulse or a spin-echo sequence.

Bo

P R dSo-^O

Bo

M t

90°

\\\\

180°

A

90°

\

t

180°

AFigure 5. Typical field cycle (idealized) for the determination of T2 by Hahn's spin-echo method. The lower plot shows the timedevelopment of the transverse magnetization Mt. Carr-Purcell sequences can be used with analogous field cycles.

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

202 Bulletin of Magnetic Resonance

Page 9: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

Zeeman levels. Alternatively, one can keep the tempera-ture constant and vary the frequency "window" to moni-tor the intensity function (Figure 6).

The second method, using field cycling, is preferablein many cases because it avoids the limitations or com-plications introduced by phase transitions and unknowntemperature dependences of the parameters. Variousactivation laws that show such complications arereviewed elsewhere (66).

2. Relaxation AnisotropyAnisotropic systems show anisotropic motions and

therefore anisotropy in the relaxation times, which canbe detected in oriented samples. This fact offers thepossibility of distinguishing different types of motions,especially different types of defects. (Figure 7 shows ani-sotropy curves calculated for dipolar interaction in al-kane chains.) Thus, relaxation anisotropy providesstructural information to add to the kinetic featuresstudied by relaxation dispersion.

According to equations 2-6, the anisotropic dipolarrelaxation rate can be written as

J (w)/ \

=-11

+ C2 <T2) (16)

where

and

a-t and <J2 a r e the mean square deviations of the dipolar

interaction functions as in equation (6).

We now define the orientation angle 0 as the anglebetween the external flux density Bo and the axis of themolecular motion. For polymers, this axis is identicalwith the chain axis. If the molecules are incompletelyoriented, the angle /3 refers to the symmetry axis of theorientation distribution. We then obtain the ratio

C2 o-2(0)

770)(17)

which depends only on C2. At fixed frequencies andtemperatures, C2 is a constant, weighting the contribu-tions of ff! and o2. The range of Cz for Debye-functions(equation 10b) is limited to 0.25 for O>TC » 1 and to 1 forWTC < 1. This means that different relaxation aniso-tropies result for different frequencies. As shown inFigure 7, the strongest anisotropy holds for C2 = 1 or o>rc

«. 1. In other words, the experimental investigation ofthe anisotropy is easier at the low frequencies accessibleby the field-cycling technique.

Figure 6. The intensity function /(o) can be scanned point bypoint by varying theZeeman-field splitting h y Bo.

3. Quadrupolar DipCross-relaxation (68,69) from dipole to quadrupole

nuclei leads to a quadrupolar dip in the ^-dispersioncurves (13,14,70,71). The reason is a crossing of theZeeman and quadrupole splittings as shown in Figure 8.The frequency range in which this effect occurs dependson the quadrupole splitting. The corresponding frequen-cies are often only accessible by the field-cyclingtechnique.

The shape of the quadrupolar dip reveals the spec-trum of the quadrupole splittings and, by circumstance,the intensity function of the interaction with quadrupolenuclei (13). The determination of the cross-relaxationrates from the quadrupolar dip allows the study of thespecific interactions between the dipole and the quad-rupole nuclei.

The equivalent effect of "spin mixing in the laboratoryframe" is also useful in a double-resonance techniquefor the detection of pure nuclear quadrupole resonances(72-74). However, this type of application of field cyclingis beyond the scope of this article.

B. Longitudinal Relaxation

The standard field-cycling program for the measure-ment of the longitudinal relaxation time r, is shown inFigure 4. The sample is polarized in as high a field aspossible. After reaching equilibrium, the field is rapidly

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

Vol. 1,No. 4 203

Page 10: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

20 40 60 80 0

oscillation jumping kinkrotation jumping kink*rotation

Figure 7. Orientation dependence of in-tramolecular mean square deviations of thedipolar interaction functions F,m<

1> and F/m<2>of the N proton pairs with numerical indices /and m in an isolated alkane chain. /} is theangle between the chain axis and the fluxdensity. Calculation includes the effect of thesixth neighbor methylene group (omission ofinteraction with a neighbor group increasesthe apparent anisotropy). Models refer towhole-chain motion around the chain axis ordiffusion of local defects along the chain.Kink defects (gtg sequences) have an as-sumed concentration of 0.15 kinks permethylene group. "Oscillation" around thechain axis is assumed to be classical with anamplitude of 18.5°. The curves show thedifferent efficiencies and anisotropies of thefunctions (67). Also see Figures 13a and 13b.

switched to the variable relaxation field, leading to anon-equilibrium magnetization, provided that the fieldwas altered fast enough (see section III. D). After a vari-able period T, the field is switched to the detection fieldaccording to the resonance condition u> = y Bo. Afterreaching resonance, the magnetization is detected by a90° pulse or an echo sequence. Alternatively, a signalcan be obtained by a fast adiabatic passage of theresonance field. The evaluation is performed accordingto

Mz(r) - = [MM - (18)

/W2(°°) is the Curie magnetization in the relaxation field,MZ(Q) is the non-equilibrium magnetization at the begin-ning of the period in the relaxation field, and 7"., is thelongitudinal relaxation time in the relaxation field.

For higher relaxation fields, the difference from thepolarization field may be too low to cause a sufficientdeviation from equilibrium. In this case, a 90° pulsewhich sets Mz = 0 immediately before switching to therelaxation field is useful. The evaluation formula is thenidentical to that of the conventional 90 ° /90 ° method (16)with Mz = 0.

In appropriate systems, the polarization of the samplecan also be enhanced by taking advantage of any kind ofdynamic polarization (2). A considerable increase in sen-sitivity can be expected (75).

C. Transverse Relaxation

The method of Hahn as well as the Carr-Purcell-Meiboom-Gill sequence (16) can be combinedwith field cycling. Figure 5 shows the field program ap-propriate for Hahn's method. The possibility of exten-sion for a Carr-Purcell sequence is obvious. The decay ofthe echoes directly yields the time constant T2 as a func-tion of the relaxation field.

D. Experimental Requirements and Limits

The field programs shown in Figures 4 and 5 areidealized. In reality, finite switching times between polar-ization, relaxation and detection fields, respectively, areunavoidable. The problem is to find the limits of theswitch intervals that will allow one to obtain reliablemeasurements.

In some field-cycling experiments described in theliterature, the direction of the external field is changedduring the field program. Then the transitions have to beslow enough to satisfy the adiabatic condition (76,77)

(19)

In all cases, where the direction of B remains stationary,this condition becomes trivial. Switching to zero externalfield Bo, however, means that B approaches the localfield BL with no definite direction. A dipolar order is

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

204 Bulletin of Magnetic Resonance

Page 11: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

(1)

= V\ B o

QX = ) X B °

Bo/T

Figure 8. Frequency survey for 35pl and 1H resonance crossing. u>A is the Uarmor frequency of 1H; yx is thegyromagnetic ratio of 35CI; OJXQ is the zero-field and cos the low-field resonance frequency of 35CI. The crosshatchedarea indicates the entire range for powder samples. (71).

produced and the condition given by equation (19) isagain relevant (77).

On the other hand, the transitions between the diversefields have to be fast enough to avoid equilibration dur-ing the switch intervals. This is certainly the case in the"fast transition limit" of the switch time AT

AT « 7"i,2(Sor) (20)

where, as the strongest restriction, the relaxation timesin the relaxation field Bo

r have been chosen. In reality,relaxation times frequently rise with Bo, so that the re-striction is less severe than it may seem.

If the course of the field during the switch intervals isreproducible from field cycle to field cycle, it can beshown that the condition given by equation (20) is far toorestrictive (78). The longitudinal relaxation curve, mea-sured after the time r + AT, is given by

MZ(T + AT) - Mz{°°)= k, A/Wzexp(- T / T^BJ) (21)

where AMZ is the difference between the magnetization

after reaching the relaxation field and the Curie magne-tization Mzo

r in this field, and k^ is a constant forreproducible field transitions. There is then no systema-tic error, even for Tf < AT. The restriction is given ratherby the condition that k^AMz must be of a measurablesize, i.e. by the sensitivity requirement. For the appara-tus referred to in ref. 78, this led to the condition

TABor)^Ar/6 (22)

for samples with an overall proportionality of approxima-tely T-i Bo. AT is the total switching period from trig-gering the field rise to resonance. (Note also that the ac-tual slope of the field rise often is initially extremely steepand becomes flatter near resonance, so that the "crit-ical" region with short relaxation times is rapidlytraversed.)

A further requirement is a sufficiently large magneticflux density Bo

d of the detection field. The reliable detec-tion of solid state signals is related to the bandwidth ofthe RF apparatus. Therefore, detection frequencies of atleast 10 MHz should be used. For liquids, the bandwidth,

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

Vol. 1, No. 4 205

Page 12: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

and therefore the detection frequency, may be lower. Inthis case, detection is possible even in the earth's mag-netic field (79,80).

There are two ways to produce the field cycle. Eitherone transports the sample between positions with differ-ent magnetic flux densities, or one switches the fieldelectronically between the desired values.

The first applications of the sample-transport methodwere reported by Pound (81) and Ramsey and Pound(82). The method was then improved by moving the sam-ple between the middle of the gap of an iron magnet anda predetermined point in the fringe field (83). Com-pressed air has been used to transport the samplethrough a tube between two magnets (69,84,85). Thus,considerably shorter transport times have beenachieved than with a manual moving device (86).

A detailed description of a compressed-air machine isavailable (87) with a reported traverse time of 250 ms.Though this is certainly not the limit of the compressedair method, it shows that applications are possible onlyfor samples with relatively long relaxation times. The di-rection of Bo may differ along the path of the sample as aconsequence of field inhomogeneities, in which case theadiabatic condition equation (19) might be relevant.

These disadvantages can be avoided by an electronicfield-switching arrangement. The use of ironless magnetcoils helps to obtain reproducible fields and preventlosses and residual magnetism in the material. For asolenoid of n windings, length C, and cross section Amuch smaller than C2, the instantaneous rate of field var-iation is

Bo(t) =U{t) - l(t) R(t)

nonA (23)

where U is the applied voltage, / the current, and R theeffective ohmic resistance of the whole circuit. Thenumber of windings n can be expressed by the magneticflux density Bo

d of the detection field and the current ldneeded to produce it. The instantaneous velocity of fieldvariation is then

^ v (24)

where the solenoid volume V = A9.. From equation (24) itbecomes obvious that in order to rapidly leave thecritical low-field range (where the relaxation times areshort), one should provide a peak power ld Umm as highas available, while the ohmic resistance and the solenoidvolume should be kept as small as possible. This meansthat U(t) should be initially higher than the voltagerequired later for steady state operation.

Equation (24) also indicates that rapid decrease of theflux density requires the rapid dissipation of the fieldenergy, and, therefore, an ohmic resistance R(t) as largeas possible. In other words, the ohmic resistance should

be switchable. One example of a practical variable resis-tance is a well cooled parallel circuit of high-voltage tran-sistors. The applied voltage should be zero or, if possi-ble, even negative during the field decay. The volume ofthe solenoid should again be as small as compatible withhomogeneity requirements.

In practice, the combination of high-power pro-grammable power supplies and small solenoids leads tocooling problems. Further limiting factors are the extentof homogeneity in the sample volume and the stabilityand reproducibility of the detection field. Various sug-gested compromises are discussed below.

In the first proper field-cycling experiments (79,80,88),a high polarizing field was non-adiabatically switched offafter a certain period leaving the earth's field as the onlymagnetic field. The polarizing field was adjusted perpen-dicular to the earth's field so that the non-adiabatic tran-sition led to a free precession signal without any RF exci-tation. A similar method, but with adiabatic field transi-tions and 90°-pulse detection, has been described (89).

Considerably higher sensitivity can be achieved withhigher detection fields. An apparatus has been des-cribed (90) which allows polarization of the sample in afield ranging from 0.05 to 1 T while the signal detection isperformed at 4 MHz (i.e., 0.1 T for protons). Liquid-nitrogen-cooled copper solenoids or a superconductingcoil have been used to reduce or eliminate the ohmic re-sistance of the solenoid wires. This machine has sincebeen improved to allow automatic computer-controlledoperation (91) for which the field-cycling method is par-ticularly suited.

A relatively simple but effective field compensatingmethod has been described (92-94). Two power supplieshave been used. One non-switchable precision powersupply produces the polarization and detection field in arather large pair of Helmholtz coils, thus insuring highstability and homogeneity. The relaxation field isproduced by a second concentric pair of Helmholtz coilsof considerably smaller diameter, together with a com-mercial programmable power supply, which generate anadditional field during the relaxation period. The super-position of the two Helmholtz fields allows a partial com-pensation or addition depending on size and polarity. Asthe volume of the compensation coils is small, shortswitching times can be achieved without deterioration ofthe detection field. The inhomogeneity of the relaxationfield produced in this way is not important at higher mag-netic flux densities, but can restrict application at thelowest fields, where the total flux density becomes un-defined as a consequence of the difference in thehomogeneities of the two pairs of coils.

Several other field-cycling machines have been used(78,95). The present apparatus in our own laboratorywas designed especially for applications with solids and

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

206 Bulletin of Magnetic Resonance

Page 13: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

for samples yielding weak signals (51). Polarization anddetection fields are in the range 0.5 to 1.2 T and we useliquid-nitrogen-cooled copper coils or a superconduct-ing solenoid wound of multifilament niobium-titaniumwires. The homogeneity in the sample volume (1 to 3cm3) is better than 5 x 10-5. The stability of the detec-tion field is about 5 x 10~5. Starting from zero, a fluxdensity of 0.1 T is passed within 4 ms with the copper coiland within 2 ms with the superconducting solenoid. Asthe course of the field rise is reproducible, TA valuesdown to 5 ms can be measured provided T^ has suf-ficient field dependence. The peak power of the pulsedpower supply is about 18 kW, and the inductance of thefield solenoids ranges from 80 to 100 mH.

IV. APPLICATIONS

A. Biological Systems

1. Protein SolutionsThe most frequent application of field cycling in NMR

relaxation spectroscopy concerns protein solutions.This is simply because the variation of the temperature isrestricted to the range where the protein solution is

1OC

neither frozen nor denatured. Thus, in contrast to manyother systems, the only possible way to study the proteinmotions by relaxation spectroscopy is to vary thefrequency.

Early relaxation studies at a single frequency (96,97)showed that water protons relax faster in protein solu-tions than in pure water even for diamagnetic proteins.Relaxation dispersion measurements were therefore de-sirable to learn something about the protein propertiesthat cause this effect.

Many proteins have been studied by this method andmany interpretations have been suggested(91,92,98-108). A typical relaxation dispersion plot isshown in Figure 9. The fact that the dispersion is spreadover almost the entire frequency range (103 to 108 Hz)leads to the conclusion that the relaxation of the waterprotons somehow reflects the protein motion. All auth-ors agree to this but suggest several quite differentmodels to explain how water molecules, which normallytumble four or five orders of magnitude faster than pro-tein molecules, can reveal the motions of thesemolecules. A relevant question is: In which part of thesystem and by which interaction partners are the nucleiof water molecules relaxed?

10-

Aqueous

-

• 00 .8

0 • •

BSA

o*#8

•0 °

1 I 11 1

solutions

X

• 8

0

. 8 °

# §

1 1 1 [ i 1 1 1 1 i l l

x 8 A

0 AA A

0 H2

• H2

x D2

A

A

0

A> A

DA

L- signal

O/D2O 50%, L-signal

O,

D D2 O,

A dry

L- signal

S+L- signal

BSA, S + L-signal

104 105 10° 10' 10° 1O9 v/Hz

Figure 9. Longitudinal proton-relaxation times versus frequency for various bovine serum albumin (BSA) solutions at 0°C (100).Concentrations are 0.391 g BSA/cm3 solvent. L is the signal detected about 200 us after the 90° pulse; S + L is the signal detected10 ixs after the start of the 90° pulse. Close agreement of the data for the various systems led to the model shown schematically inFigure 10.

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

Vol. 1,No.4 207

Page 14: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

Several authors have suggested that more or lessirrotationally-bound water molecules in the hydrationshell relax the bulk water by a fast material-exchangeprocess (98,99,107), so that the rotational correlationtime of the macromolecule becomes relevant (two-siteexchange model). Koenig and Schillinger (98) havepointed out that the water molecules need not be boundrigidly on the protein surface for this argument to bevalid. Rather, it would be sufficient if the motion of thebound water was restricted to rotational diffusionaround one axis fixed on the macromolecule. Clearly,this type of model includes many variants concerningmore or less restricted and continuously or discretelydistributed degrees of motional freedom within thehydration shell. The discussion concerning all two-siteexchange models can be concentrated on two crucialpoints: the mean residence time rM of a water proton inthe bound state and the contributions of intra- and inter-molecular interaction to the water relaxation.

The mean residence time TM of the water protons andthe number of water molecules in the bound state withinthe hydration layer have been dealt with by Koenig et al(105). Assuming a two-site exchange formula, theseauthors estimated the limits of rM on the basis of 1H, 2Hand 17O relaxation data. The argument used in this es-timation was that rM should be longer than the correla-tion time relevant for the rotational diffusion of the pro-tein molecule, but shorter than the measured relaxationtimes. This follows from the fact that the protein rota-tional diffusion is visible in the ^-dispersion of the waterprotons and also from the observation of exponentialrelaxation curves characteristic of the fast exchangelimit. The main advantage of the use of 2H and 17O relax-ation data for this estimation is that these isotopes showvery short relaxation times, so that the upper limit of TM

could be reduced by two orders of magnitude.

According to Koenig et al (105), the ranges of possiblevalues of TM estimated in this way for solutions of a smallprotein (lysozyme, molecular weight 14,700) and a largeprotein (hemocyanin, molecular weight 9 x 106) are notcompatible with a common rM value or a common dis-tribution of rM values. This conclusion depends on howmuch reliance is placed on the estimated limits of rM.

Another objection against all simple two-site ex-change models concerns the order of magnitude of TM.The tentative explanation of the relaxation data by sucha model leads to values of rM five orders of magnitudelarger than is compatible with the binding energy ofhydrogen bonds (105), which could be assumed to bethe main reason for the formation of the hydration layer.On the other hand, it is well known that rM values in therequired order of magnitude occur, for instance, in neu-tral solutions of small ions (12,98). Therefore, the elec-trostatic interaction with a few charged groups on the

protein may explain such long residence times in thecase of protein solutions.

Furthermore, one has to consider exchangeable pro-tons of the protein molecule itself, so that in the case ofthe hydrogen isotopes TM has to be considered rather asan exchange time between protein and water than as aresidence time of water molecules in the hydration shell.

This explanation is supported by the observation thatthe number of protein-associated exchanging nuclei isessentially smaller than the number of water moleculesrequired to form a single-layered, closed hydration shell(105). Below, the term "exchange" will be generalized tomean non-material spin exchange, leading to a com-bined material- and magnetization-transfer model,which avoids the problems mentioned above.

If dipolar interaction among the water protons werethe only mechanism of relaxation, i.e., if the two sites as-sumed in this model were isolated with respect to inter-actions with other parts of the system, an isotopic dilu-tion experiment should show a strong lengthening of therelaxation times according to the degree of dilution.Experiments of this type (100,101) showed almost noisotopic-dilution effect for serum albumin solutions of asufficiently high concentration (Figure 9). This led to theunambiguous conclusion that the water protons arerelaxed by an interaction with spins in the protein whichare not affected by the isotopic dilution. Clearly, in thistype of investigation one must take into account thesmall but not negligible increase of the water viscosityupon deuteration (109), and a sufficiently high proteinconcentration, depending on the individual protein, isnecessary for reliable results. The conclusions forlysozyme solutions (110,111) were drawn without con-sidering these conditions.

A hydrodynamic ordering of unbound water sur-rounding the protein molecules has been suggested(91,105). The motion of the water molecules should thenbe anisotropic, leaving a small but slowly reorientingcomponent. However, the hydrodynamic water-orderingmodel again contradicts the finding that the water pro-tons are relaxed via /ntermolecular water-protein inter-action especially at higher protein concentrations.

The models mentioned so far do not take explicitlyintouaccount the protons of the protein. This oversim-plifies the system, which is in fact highly heterogeneousin the sense of section II.D. Observations with serum al-bumin indicated that the free induction decay of the pro-tein protons is composed of at least two contributions,namely a signal behaving like a solid and anothercomponent decaying like a liquid, and also that the lon-gitudinal relaxation curve is exponential, except for pro-teins like gelatin. This led to a combined material-exchange and magnetization-transfer model (100,101).Figure 10 gives a schematic representation of the model.

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

208 Bulletin of Magnetic Resonance

Page 15: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

protein

"mobile" protons

* L-signal of the protein

MT

rigid " protons

= S -signal

Dl, MT

CE

Dl, MT

V//////X.

solvent

hydration

shell

CE unbound

water

Figure 10. Schematic representation of the exchange processes that in-fluence proton relaxation in aqueous protein solutions (101). MT indicatesmagnetization transfer, Dl dipolar interaction, and CE chemical exchange.

We have used the term "magnetization transfer" in-stead of the terms "spin diffusion" or "cross relaxation"usually used for extended solids and direct interactionpartners, respectively. It is rather a matter of tastewhether one considers a more or less rigid mac-romolecule as a solidlike structure showing spin diffu-sion due to mutual flip-flop processes, or one empha-sizes the heterogeneous mobility of the protein protonsso that the relevant scheme is a complex multinucleicross-relaxation which includes more than flip-flop tran-sitions. Both views are justified in a certain sense. Theterm "magnetization transfer", thus far not associatedwith any special system, may therefore be preferable.

The essence of the model shown in Figure 10, or invariants of it, has now been adopted by several authors(108,112-118). Campbell and Freeman have treated thepopulation dynamics of coupled spins with differentrelaxation rates, due to additional specific interactions,on the basis of Solomon's theory (6). They find that crossrelaxation nevertheless can lead to the exponentialrelaxation curves most frequently observed. A similarapplication of Solomon's theory has been given by Kalkand Berendsen (114). They believe that rapidly rotatingmethyl groups act as relaxation sinks which relax thetotal protein at sufficiently high frequencies. Edzes andSamulski (113, 117) and Koenig et al (108) have givensystems of rate equations for the global solvent and pro-tein magnetizations. Thus, it seems to be well estab-lished that magnetization transfer plays an importantrole in the complex "protein water", albeit the correctanalytical description is still under discussion.

A special question concerns the existence of relaxa-tion sinks within the protein molecules. These sinksshould fulfil some essential requirements. They should

be accessible to all spins in the sense of the exchangescheme of Figure 10. Deuteron dilution should also dilutethe number of protons directly accessible to the sinks,so that the solvent relaxation is unaffected while the non-exchangeable protein protons may have even longerrelaxation times (see ref. 113 and Figure 9). Further-more, different mobilities are required (100).

As a suitable candidate for the sinks, we have sug-gested the protons in the NH groups within the protein(13,119,120). It can be shown that NH protons dispersedin a "bath" of the other protons account for the exper-imental facts. Furthermore, the quadrupole dip (sectionIII.A.3) due to NH cross relaxation has been observed indry bovine serum albumin, directly verifying this type ofrelaxation sink at low frequencies (121). Of course, theremay be competing sinks depending on the system andthe experimental conditions, for example, the paramag-netic heme group in hemoglobin (96,102) or methylgroups at high frequencies (114)!

The explanation of the special frequency dependenceof r, observed in protein solutions is difficult. As the pro-tein molecule contains the sinks of the total relaxation,one has to deal with a heterogeneous system not de-scribable by a single correlation function. At least twodifferently relaxing components occur, as is revealed bythe wide-band shape of the protein-free induction decay(100). The problem then is that the type of the individualcorrelation function is known only for groups sufferingthe overall tumbling of the molecule alone, but not forthose showing internal motions, if no special assump-tions are made (see section II.C). Furthermore, the in-fluence of spin diffusion itself on r, dispersion must beconsidered (101).

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

Vol. 1,No. 4 209

Page 16: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

Nevertheless, information can be derived concerningthe correlation time of the rotational diffusion of thewhole molecule and, with less assurance, of the motionsof the flexible groups. For example, the former correla-tion time could clearly be related to the molecular weightof the protein (91). Thus, an empirical distribution of cor-relation functions has been used simply to get a quanti-tative description of the experimental data. In the samesense, the intermolecular interaction between hemo-globin molecules at very high concentrations could bestudied (104,106).

The relaxation enhancement of water protons by par-amagnetic ions or radicals bound to enzymes or othermacromolecules is an appropriate tool for studyingbinding sites (122-133). The relaxation enhancementupon addition of macromolecules to a solution of par-amagnetic ions or radicals is due to the different relaxa-tion parameters in the bound state of the paramagneticsubstances. Most frequently, Mn2+ has been used,which interacts with the surrounding protons by dipolarand scalar coupling (Table 1). A number of parametersdetermine these interactions because the electron relax-ation time must be taken into account as a function ofthe magnetic flux density. It was therefore necessary toinvestigate a large frequency range by the field-cyclingmethod to allow less ambiguous fits of all parameters.Examples have been discussed (134-138) in terms of theproton exchange behavior between the paramagneticcenter and the solution; the exchange behavior of theparamagnetic ion between the binding sites on the pro-tein and the free water; the electron-proton distance;and the correlation times for the dipolar and, if relevant,scalar interaction. This is treated in section IV.D in con-nection with solutions of paramagnetic ions.

2. Model MembranesLipid bilayers are model structures for biological

membranes (139). The type of molecular motion in thesemodel membranes is of special interest with respect tothe permeability properties of membranes. Oneproposed model for passive transport assumes diffusingdefects of the lipid structure to be "carriers" for smallpermeants (140). In an alternative model, the diffusion ofvacancies provides the free volume required for a diffu-sion step of the permeant. To test these views, it isnecessary to verify the special type of fluctuation con-nected with defect diffusion (section II.C). Apart fromsuch experimental tests, general investigations of flex-ibility help to characterize and provide a physical under-standing of the different thermodynamic phases ofmembranes.

Many papers (141-144) treat molecular dynamics inmodel membranes. Field-cycling measurements in boththe gel and crystalline phase of dipalmitoyl-lecithinbilayers have been reported (145). Two frequencyregions in the 7, dispersion can be distinguished (Figure11). The low-frequency process, clearly visible in theliquid-crystalline state, must be attributed to coopera-tive motions of many chains. Such "order fluctuations"will be discussed in section IV.B.

The high-frequency process has been interpreted interms of defect diffusion limited by two reflecting bar-riers, namely the polar headgroups terminating the lipidmolecules (23,51; also see section II.C and Figure 3).Figure 11 shows the comparison between theory andexperiment. It was assumed that the transition from thegel to the liquid crystalline state affects only the pre-exponential factor of the Arrhenius-activation law for thedefect motion. Another simplification was the neglect of

1s

10.-2

10"

40% DPL in D2O

104 105 10e- i — i i i n i l .

10' v/Hz

Figure 11. Frequency dependence of the longitudinal proton-relaxation time Ti of a dipalmitoyl lecithin (DPL) aqueousdispersion (143). Phase was a gel at 35°C, a liquid crystal at60°C. Solid lines indicate calculations for the limited defect-diffusion model (equation 15 and Figure 3) with r^Tb = 227.The pre-exponential factor of the Arrhenius activation law fordefect-diffusion times was assumed to decrease by a factor0.35 at the gel-liquid crystal transition. The dashed line repre-sents TJTO =113 and a ratio 0.25 of the pre-exponential fac-tors. Arrows indicate the positions wTd = 1. Order fluctuationsare believed to cause the decay at low frequencies in the liquidcrystalline phase.

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

210 Bulletin of Magnetic Resonance

Page 17: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

the so-called flexibility gradient along the alkyl chains ofthe lipids (37,143). An attempt has been made to takethis gradient into account (146), using the numericalMarkov chain method (23) and assuming that the diffu-sion step rate varies along the alkyl chains while thedefect concentration is constant. The result is that thedispersion of the non-selective proton-relaxation timesis almost the same as expected in the case of constantdefect-diffusion step rates.

B. Cells and Tissues

The relaxation times of cell and tissue water is re-markably sensitive to alterations of the cell state. Forexample, it has been found that water relaxation timesare affected by mitosis (147), tissue malignancy(148,149), and ischemic damage (149,150). Even theconversion of the spin state of a single protein (cyto-chrome P-450 complexes) influences water relaxationtimes (151). The relevant relaxation mechanisms aretherefore of considerable interest.

Unfortunately, cells and tissues are extremely heter-ogeneous materials. Apart from the problems men-tioned in the previous sections in connection with pro-tein solutions and membranes, one has to distinguishbetween intra- and extracellular water. Several, some-what controversial papers deal especially with heter-ogeneity in the water phase and the exchange mecha-nisms (117,152-158). In general, one can state that thecell proteins are the centers responsible for the totalrelaxation behavior.

Reported 7Vdispersion data for erythrocytes andwhole blood (101,103) confirm this statement. Thesesources also concluded that encapsulation of hemo-globin within cells has little effect on its molecular mo-tion, while oxygenated sickle cell hemoglobin forms ag-gregates even in cell-free solutions, as indicated by amore hindered molecular tumbling.

Held et al (159) have measured the r, dispersion ofmuscle water over four decades. The fact that no finallow-frequency plateau could be observed in the avarl-able frequency range can be explained by the large pro-tein aggregates within muscle cells.

Fung (160) has also investigated the frequency depen-dence of Tf for protons and deuterons in muscle water.An important result is that (as in solutions of globularproteins) isotopic dilution of cell water by D2O has al-most no effect on the proton relaxation rates. This is aclear indication that the relaxation sinks are not locatedin the water phase. Probably, a relaxation schemeanalogous to that shown in Figure 10 is relevant (101).The magnetization transfer from water protons to pro-tein protons can also be deduced from the fact that

7"1(2H)/ Ti(nO) in muscle water is close to that in pure

water, while f ^ H ) / F ^ O ) is 2.1 times greater in purewater than in muscle water (153). Magnetization transferrequires dipolar or scalar interaction between the par-ticipating nuclei as a dominant relaxation mechanism. 2Hand 17O, however, are quadrupole nuclei relaxed mainlyby the interaction with electric field gradients (Table 1).

C. Liquid Crystals

7"rdispersion measurements by the field-cyclingtechnique play a crucial role in the physics of liquidcrystals. As molecular dynamics in liquid crystals areanisotropic, differences in the fluctuation rates aroundthe different axes of the molecules are expected. Orderfluctuations of the local director are of special interestwith respect to liquid crystalline properties (161-163).The problem is experimental distinction between thediverse components of molecular motion.

As order fluctuations and rotational or translationaldiffusion (45,165) of the molecules are relevant at differ-ent frequencies, it may be possible to use the frequencydependence of 7", to distinguish between the variouscomponents of motion. Experimental investigations notusing the field-cycling technique have been described(164-166). The relaxation behavior of liquid crystallinesolutions of a polypeptide (poly-7-benzyl-L-glutamate)has been reported (101,167), but the frequency range ofthe conventional technique used was not adequate for adecisive analysis of all components of molecular motion.

Noack and co-workers performed field-cycling mea-surements of p-azoxyanisole and p-methoxybenzyli-dene-p-butylaniline (168,169) and interpreted their re-sults at low frequencies (about 104 Hz) according to thetheory of order fluctuations (161). For the 7"-, dispersionat higher frequencies, they assumed isotropic transla-tional diffusion (see equations 11a, b) and anisotropicrotational diffusion (19). In their "fast exchange" formula(ref. 169, equation 1a), the contributions of different pro-ton phases are mixed with contributions of differentcomponents of motion in the same proton phase. Thelatter can occur for motions separately affecting intra-and intermolecular interactions, i.e., different protonpairs. For a particular proton pair, however, differentcomponents of molecular motion must be combined onthe level of the correlation function. Otherwise, the resultis valid only under the condition that the correlationtimes are sufficiently different from each other. Theproblem can be partially circumvented by studyingseparately the intra- and intermolecular contributionsand the motion of definite groups with specificallydeuterated compounds. Thus, the ring protons in p-azoxyanisole have been investigated at several frequen-cies with the use of partially and perdeuterated

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

Vol. 1.NO.4 211

Page 18: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

molecules (166). From this, one may conclude that field-cycling data specified in this sense would help to sepa-rate the different components so that a final picturecould be obtained of molecular motions in liquidcrystals.

A useful parameter for the distinction between differ-ent types of molecular motion is the angle between thedirector and the external magnetic field (see sectionIII.A.2). By this means, the anisotropy of the longitudinalrelaxation time 7", has been investigated experimentally(170) and theoretically (45).

Field-cycling data were reported for the liquid-crystalline phase of aqueous dispersions of dipalmitoyl-lecithin, (see Figure 11 and ref. 145). Though themolecular structure of this substance is characterized bylong fatty-acid residues and differs therefore from theclass of substances usually referred to in connectionwith liquid crystals, the ^-dispersion was quite similarto that found for p-methoxybenzylidene-p-butylaniline(169). This fact indicates that the low-frequency disper-sion is again influenced by order fluctuations. The upper-frequency range was assumed to be dominated bystructural defects diffusing up and down the alkyl chainsas described in section IV.A.2. The frequency depen-dence of 7 caused by this process contradicts the "ex-treme narrowing" case (2) occasionally assumed for thisfrequency-temperature range (171).

D. Synthetic Oligomers and Polymers

1. Solutions and MeltsThe problems investigated with synthetic polymer

solutions are similar but less complicated than with pro-tein solutions (section IV.A. 1). Again, questions such asexchange kinetics and interactions in the solvation shell(172-174) and the mobility and conformationai changesof the macromolecules themselves (175-180) are ofinterest. The latter is also the central subject studiedwith polymer melts (181,182). It is clear that special em-phasis must be laid on the anisotropy of moleculardynamics in these systems.

Both solutions and melts of polyethylene oxide havebeen studied with the field-cycling technique (183). Itwas shown that melts and solutions underlie differenttypes of processes of molecular motion. These arerevealed at low frequencies where a pronounced depen-dence of the relaxation times on the chain length was ob-served. The data were interpreted in terms of the trans-lational diffusion model for intermolecular interaction(equation 11). However, such an interpretation is doubt-ful because intermolecular interaction is of minor impor-tance compared with the intrachain contributions andhas no influence at low frequencies. This was demon-strated for polyethylene melts by dissolving a few un-

deuterated chains in a perdeuterated matrix (184,185)so that the intermolecular interaction was suppressed.As a result, the relaxation can be understood by thesame processes as in the undiluted case.

A different interpretation has been given with thethree-component model of molecular dynamics inpolymer melts (186,187). The components are aniso-tropic segment reorientation by defect rearrangements,longitudinal chain diffusion due to defect diffusion, andthe conformationai fluctuation of the environmental tubeincorporating the reference chain. The latter can be in-terpreted as a consequence of chain diffusion, so that allcomponents are deduced from elementary processesaccessible to thermal activation. The model successfullydescribes the molecular weight and frequency depen-dences of the relaxation times (Figure 12) and also ver-ifies predictions concerning the behavior of mixtures ofpolymers with different chain lengths (184,185). Therelaxation formula used in these papers is an extensionof equation (16) from the longitudinal chain diffusionprocess alone to the case where all three componentsare relevant. The extended correlation function has alsobeen used to calculate the shape of the free inductiondecays according to equation (8) (188).

For sufficiently high molecular weights and lowfrequencies, the contribution of the longitudinal chaindiffusion process dominates, and the simpler equation(15) can be used. This has been done to derive informa-tion about the correlation length of the environmentaltube surrounding the reference chain (37). For this pur-pose, data of the self-diffusion coefficient have beencombined with the characteristic time constant for thelongitudinal chain-diffusion. A similar treatment hasbeen given for polyethylene (185). The close relationshipof nuclear magnetic relaxation in polymer melts withself-diffusion has also been demonstrated by a com-puter simulation of the model (48).

From a different point of view, namely by consideringconformationai jumps of a chain described in a tetrahe-dral lattice, Valeur et al (189) obtained the same expres-sion given in equation (15a) (also see ref. 190). However,the parameters introduced in these calculations have adifferent meaning. As no interchain effects are included,these models as well as Ullman's treatment, based onthe Zimm-Rouse theory, (179,180) are relevant forpolymers in solutions. Unfortunately, no comparisonwith field-cycling data has been reported so far. Such acomparison would be a valuable test of the models,especially for long chain lengths.

2. SolidsPolymers and oligomers in the solid state are either

crystalline, partially crystalline, or amorphous. The com-mon feature of the relaxation processes in all three

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

212 Bulletin of Magnetic Resonance

Page 19: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

Figure 12. Frequency dependence ofthe longitudinal proton-relaxation time7", of polyethylene oxide melts at 70 °C.M is the molecular weight. The data(183) closely fit the solid lines derivedfrom the three-component model forpolymer melts (186,187).

T,sec

»"'•

10-

M=1OOO

M=4OOO

0=7OC

M=27OOO

10' io7

_ J _rad/sec

states is that they are local; the chain length has no di-rect influence. The anisotropy of molecular dynamics isagain of central interest. Models assuming any kind of"molecular tumbling" are questionable in this light. In-stead, local segment reorientation, assumed to result inpart from the diffusion of chain defects (for example, gtgsequences or torsions), is the proper starting point fordiscussing these problems, as was realized in earlypapers (191-195).

The development of an improved field-cycling tech-nique for solids with short relaxation times (78,95) al-lowed an experimental test of the defect-diffusionmodels discussed in section II.C. The 7"i dispersion ofpartially crystalline polyethylene has been investigatedin the ranges 104 to 108 Hz and 77 to 373 K (51,78,196).Apart from end-group rotation, which becomes visible atlower molecular weights, two discrete chain processesattributable to the amorphous regions were detected.The homogeneity of the amorphous regions with respectto nuclear magnetic relaxation was demonstrated by anisotopical dilution experiment in which spin diffusion isinterrupted. Taking advantage of the relaxation en-hancement by (paramagnetic) oxygen, which is dis-solved only in the amorphous regions even at rather highpressures, it was shown that the amorphous parts vir-tually behave as a homogeneous phase (51). This meansthat a single correlation function holds for the amor-phous methylene groups and that there is no distributionof correlation functions (section II.D).

The assumption that methyl-group rotation is initiatedby local density fluctuations as a consequence ofvacancy diffusion led to the suggestion that three-dimensional diffusion of defects with restoring capability(equation 14), or one-dimensional diffusion of defectswithout restoring capability (equation 12), might be

relevant for the kinetics of this process. In fact, these ex-pressions describe the T-, dispersion at liquid-nitrogentemperature where methyl-group rotation is dominant,at least for short chain lengths. Alternatively, a Cole-Davidson distribution of correlation times can be as-sumed (51,78). In this case, the dilution experiment de-scribed above could not be carried out because no ap-propriate samples were available.

At about - 50° C, the so-called y process dominatesthe T-, dispersion. Assuming diffusing kinks (gtgsequences) reflected at each other, this process couldbe explained by equation (15) for the limited one-dimensional defect-diffusion case (196). At highertemperatures (about 30° C), the so-called /? processfinally becomes relevant. This process was also thoughtto originate from limited one-dimensional defect diffu-sion, but, in the presence of a second type of defect, waspresumed to diffuse independently between the crystal-lites which terminate at both sides of the amorphousparts of the chain. Extended chain torsions may be anappropriate example of such a defect (196).

Field cycling is suitable for studying the physical na-ture of the defects in addition to their kinetic properties.As outlined in section III.A.2, the anisotropy of nuclearmagnetic relaxation, apparently enhanced at lowfrequencies, is a potential tool for investigating thestructural rearrangements which cause the relaxation,^-anisotropy investigations found only a small effect,probably because the experimental frequency (30 MHz)was too high (197). Our group has performed the firstexperiments (unpublished) at frequencies down to 104

Hz on polyethylene and polymethylmethacrylate with theaim of distinguishing among the various models present-ed in Figure 13. Though the broad distribution of thechain orientations has not yet allowed us to make a final

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

Vol. 1,No. 4 213

Page 20: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

T,(3)Tito)

1.7

1.4

1.3

1.2

1.1

1 .0 •

Q9

0.8

1.7

i 1-6

1.5

1.4

1.3

1.2

1.1

1.0

0.9

0.8

oscillatton

0.25

20 40 60 80 P

jumping kink

T,(P) •Ti(0)

2.4

2.2

2.0-

1.8

1.6

1.4

1.2-

1.0-

1.7-,

T,(B) r

1.4

1.3-

1.2

1.1

1.0-

0.9

0.8

rotation

0.25

20 40 60 80 P

jumping kinkrotation

1,0-

T,(P) 0.9T,(0)

0,8-

0,7-

0,6-

0,5

0,4-

0,3-

0,2-

flip-flop

0 20 40 60 80 0 20 40 60 80 P 20 40 60 80 B

Figure 13. Orientation dependence of the anisotropy quotient T-,(fi)/ 7,(0) for different kinetic models, (a) Molecular motion in al-kane chains (see Figure 7 caption). Solid lines represent the intramolecular contribution; small, dotted segments indicate com-bined intra- and crystalline intermolecular contributions, (b) 180° flip-flop motions in an alkane crystal. Data include combinedintra- and intermolecular contributions in a crystalline arrangement. Curves are labeled with C2 values. (67)

decision, the superposition of diffusing kinks and tor-sions does seem to be compatible with the T, anisotropyof polyethylene at room temperature (67). In polymethyl-methacrylate, methyl-group rotation is a dominantprocess even at room temperature (198). A 7, anisot-ropy of a factor 2 was observed below 105 Hz (199).

Crystalline paraffins have been studied by the field-cycling technique (95). A more-or-less hindered end-group rotation was characterized. The methylene groupdynamics in paraffins is of special interest because, incontrast to polyethylene, it occurs here in a pure crystal-line phase. Certain mobility changes appear at the so-called rotator transition, a solid state phase transitionbelow the melting point. Unfortunately, no information

was given about the history of the samples.Samples crystallized from the melt do not necessarily

behave exactly like samples crystallized from solution(146). It is still not clear whether the processes observedin the crystalline phase originate from a homogeneouslyrelaxing system or from local perturbations, for exam-ple, at the inner grain surfaces characteristic of melt-crystallized material. The former was assumed by theauthors in their use of a defect-diffusion model to de-scribe the r, dispersion. The Hunt and Powles expres-sion (47) has also been applied, but, as mentioned insection II.C, the assumptions for this model are some-what inconsistent. Reconsideration of modified defect-diffusion models is desirable.

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

214 Bulletin of Magnetic Resonance

I

Page 21: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

E. Non-macromolecular Systems withElectron-paramagnetic Centers

The presence of paramagnetic centers often causesdominating enhancement of the nuclear magnetic relax-ation rates as a consequence of dipolar and scalar inter-actions with these centers (Table 1). The investigation ofthis frequency-dependent enhancement allows conclu-sions concerning the properties and dynamics of theseinteractions. Examples for such centers are paramag-netic ions (12), free radicals (200), and excitons (201). Ashort review of applications in connection with enzymeswas given in section IV.A.1. In the following, we considerparticularly systems of small molecules suitable forstudying the basic interactions.

Manganese ions in solution are an example of asystem in which scalar interaction is revealed by a low-frequency step in the 7", dispersion. In addition to thecoupling constant, the exchange time in the solvationsphere and/or the electronic relaxation time can be in-vestigated. The latter is possible for scalar interaction, incontrast to dipolar coupling, because the correlationtime for scalar interaction depends solely on the relaxa-tion time of the unpaired electron and the exchange timeof the solvation shell. If the electronic relaxation time islong enough, the exchange time can even dominate thescalar correlation time (11,12).

Field cycling has been applied to the study of manga-nese solutions (88,89,93). The transverse relaxation timer2 has also been investigated with this technique, using aCarr-Purcell-Gill-Meiboom pulse sequence (ref. 93 andFigure 14). The use of different solvents in these studiesmodified the parameters of the solvation complex.

An example in which scalar interaction plays no role issolutions of Gd3 + (202). The exchange behavior of thesolvation sphere is then not generally deducible fromr rdispersion data, i.e., from the height of the dispersionstep at WSTC = 1, where cos is the Larmor frequency of theunpaired electron and rc is the effective correlation timeof the dipolar interaction. This step height is reduced inthe so-called hindered exchange case TM <« T1M, whereTM is the mean residence time of a solvent proton andT1M is the longitudinal proton relaxation time in the firstsolvation sphere (203,204). The same effect can also bederived from an assumption of a field-dependent cor-relation time, i.e., electronic relaxation times shortenough to be comparable with the rotational correlationtime of the solvation complex. An extensive discussionof this ambiguity is given in ref. 202.

In the last few years, nitroxyl radicals have been ofspecial interest because they are used to spin labelmacromolecules (133). Such radicals have been studiedin various solvents with the field-cycling technique(205,206). Scalar interaction is again present in these

10 -

10

10 -

Figure 14. Frequency dependence of the molar proton-relaxation rates cT^ -1(0) and cT2~ \H) of a solution of Mn(CI04)2in dimethylsulfoxide (DMSO) at various temperatures. Thecontribution of bulk-phase DMSO has been subtracted. (93,reprinted by permission).

systems. A complete set of structural and kinetic pa-rameters of the solvation complexes was derived.

Excitons in molecular crystals are a quite differenttype of paramagnetic center. Naphthalene and anthra-cene crystals irradiated with light have been investigated(201,207,208). Because the non-paramagnetic contribu-tion to the relaxation rate is rather weak, the effectiverelaxation time is given solely by the exciton-proton

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

Vol. 1,No. 4 215

Page 22: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

interaction. In these investigations, the relaxation timeswere so long that the field was varied by manual up-and-down switching. Information concerning exciton mi-gration via a hopping mechanism could thus be derived.

We finish this review with this application of the field-cycling technique to a system which could also be con-sidered a subject of solid state physics. Further applica-tions of solid state physics are considered beyond thepresent scope. It must be emphasized, however, that im-portant studies with this method have been performed inthat field too. Applications to superconductors couldeven be considered as crucial experiments for Cooperpairing and other phenomena (209-211). Here one ex-ploits the possibility of detecting NMR signals in the nor-mal state by field cycling across the critical fieldstrength.

V. FINAL REMARKS

The aim of this review article was to outline thecapabilities of the field-cycling technique in nuclearmagnetic relaxation spectroscopy. At the present stateof the art, the method allows the investigation of thefrequency dependence of relaxation times of liquids andsolids in a broad range (103 to 108 Hz for protons) butwithout resolution into individual resonance lines. Sever-al applications are reported which have clarified am-biguities left by conventional relaxation studies. On theother hand, studies would be more detailed if the exper-iments could be carried out under high-resolution condi-tions. Whether this will ever be possible is doubtfulbecause of extreme technical difficulties. The mostproblematic requirement for such experiments is thegeneration of homogeneous and stable detection fieldswithin a few milliseconds. Thus, it cannot be expectedthat the resolution of the method will be greatly im-proved in the near future. Rather, the sensitivity will in-crease so that the technique will be applicable tochemically or isotopically labeled substances, i.e., to lowconcentrations of the resonant nuclei.

The field-cycling technique is most frequently appliedto the study of molecular dynamics. Non-NMR methods,for example, quasi-elastic light or neutron scattering anddielectric relaxation, are available for the detection orderivation of correlation and intensity functions. Thesemethods can have advantages, such as lower cost ormore desirable frequency ranges. The main advantagesof the NMR field-cycling technique are its applicability toalmost all kinds of systems and a straightforward con-nection between theory and experiment worked out formany models. Field cycling can thus be considered afavorable compromise with respect to frequency band-width and broad range of applications.

References

1 N. Bioembergen, E.M. Purcell, and R.V. Pound, Phys. Rev. 73,679(1948).

2 A. Abragam, The Principles of Nuclear Magnetism, Oxford Univer-sity Press, Oxford, 1961.

3C.P. Slichter, Principles of Magnetic Resonance, 2nd ed.,Springer-Verlag, Berlin, 1978.

4 M. Goldman, Spin Temperature and Nuclear Magnetic Resonancein Solids, Oxford University Press, Oxford, 1970, chapters 3 and 6.

5D.C. Ailion, Adv. Magn. Reson.5,177(1971).6 1 . Solomon, Phys. Rev. 99,559 (1955).7 P.S. Hubbard, Phys. Rev. 128,650 (1962).8 J . Schriever and J.C. Leyte, in Proceedings of the European

Conference on NMR of Macromolecules, Sassari, 1978, F. Conti, Ed.,Lerici Edizioni, Rome, 1978.

9P.W. Anderson and P.R. Weiss, Rev. Modern Phys.25,269(1953).10 P.W. Anderson, J. Phys. Soc. Jpn. 9,316 (1954).11 L.O. Morgan and A.W. Nolle, J. Chem. Phys. 31,365 (1959).12 R. Hausser and F. Noack, Z Phys. 182,93 (1964).13 R. Kimmich, Z Naturforsch. Teil A 32,544 (1977).14 H.T. Stokes and D.C. Ailion, J. Chem. Phys. 70,3572 (1979).15G.C. Levy, D.M. White, and F.A.L. Anet, J. Magn. Reson. 6, 453

(1972).16T.C. Farrar and E.D. Becker, Pulse and Fourier Transform NMR,

Academic Press, New York and London, 1971.17A.A. Maryott, T.C. Farrar, and M.S. Malmberg, J. Chem. Phys. 54,

64(1971).18W.B. Moniz and H.S. Gutowsky, J. Chem. Phys. 38,1155 (1963).19 D.E. Woessner, J. Chem. Phys. 37,647 (1962).2 0N. Bioembergen, Phys. Rev. 104,1542 (1956).2 ' R. Kimmich, Colloid & Polym. Sci. 252,786 (1974); 254,918 (1976).22 H.C. Torrey, Phys. Rev. 92,962 (1953).23 R. Kimmich and A. Peters, J. Magn. Reson. 19,144 (1975).24 R. Kimmich and G. Voight, Z Naturforsch. Teil A 33,1294 (1978).25 M. Eisenstadt and A .G . Redf ie ld , Phys. Rev. 132,635 (1963).26 D. Wol f , J. Magn. Reson. 17 ,1 (1975).27 P. Bordewijk, Chem. Phys. Z.ert.32,592(1975).28 R. K immich , Polymer^, 851 (1975).29 P. Debye, Polar Molecules, Dover, New York , 1945.3 0 F . Noack , G.J. Kr i iger , W. Mu l le r -Warmuth , and R. Van S teen -

winkel, Z Naturforsch. Teil A 22,2102 (1967).3 1U. Haeberlen and G. Maier, Z Naturforsch. Teil A 22,1236 (1967).32 D.C. Look and I.J. Lowe, J. Chem. Phys. 44,2995 (1966).33E.R. Andrew, R. Gaspar, T.J. Green, and W. Vennart, in Magnetic

Resonance and Related Phenomena, H. Brunner, K.H. Hausser, and D.Schweitzer, Eds., Groupement Ampere, Heidelberg-Geneva, 1976, p.131.

34S.H. Koenig, BiopolymersU, 2421 (1975).35D.E. Woessner, J. Chem. Phys. 36,1 (1962).36D.E. Woessner, B.S. Snowden, and G.H. Meyer, J. Chem. Phys. 50,

719(1969).37Y.K. Levine, N.J.M. Birdsall, A.G. Lee, J.C. Metcalfe, P. Partington,

andG.C.K. Roberts, J. Chem. Phys. 60,2890 (1974).38 H. Si l lescu, Advan. Mol. Relaxation Processes^, 91 (1972).39 D. Beckert and H. Pfeifer, Ann. Phys. (Leipzig) 16,262 (1965).40F.-J. Fritz, B. Huppert, and W. Willems, Stochastische Matrizen,

Springer-Verlag, Berlin, Heidelberg, and New York, 1979.4'J.E.Anderson and R. Ullman, J. Chem. P/iys.47,2178(1967).42 H. Pfeifer, Ann. Phys. (Le ipz ig )8 ,1 (1961).43 J.F. Harmon and B.H. Muller, Phys. Rev. 182,400 (1969).44 G. Held and F. Noack, in Magnetic Resonance and Related Phen-

omena, E.R. Andrew and C.A. Bates, Eds., North Holland, Amsterdam,1975, p. 461.

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

216 Bulletin of Magnetic Resonance

Page 23: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

45S. Zumer and M. Vilfan, Phys. Rev. A17,424(1978).46S.H. Glarum, J. Chem. Phys.33,639(1960).47 B.I. Hunt and J.G. Powles, Proa Phys. Soc. LondonM, 513 (1966).48 R. Kimmich and W. Doster, J. Polym. Sci., Polym. Phys. ed., 14,

1671(1976).49Y.K. Levine, J. Magn. Reson. 11,421 (1973).50 R. Kimmich, Z Naturforsch. Teil A 31,693 (1976).51C.V. Heer, Statistical Mechanics, Kinetic Theory, and Stochastic

Processes, Academic Press, New York and London, 1972.52 R. Orbach, in Fluctuation, Relaxation and Resonance in Magnetic

Systems, D. Ter Haar, Ed., Oliver and Boyd, Edinburgh and London,1962.

53 D. Shaw, Fourier Transform NMR-Spectroscopy, Elsevier, Am-sterdam, Oxford, and New York, 1976.

54 F. Noack and G. Preissing, Z Naturforsch. Teil A 24,143 (1969).55H.E. Rorschach, P/?ys/ca30,38(1964).56G.R. Khutsishvili, in Magnetic Resonance and Related Phenom-

ena, V. Hovi, Ed., North Holland, Amsterdam and London, 1973, p. 35.57 U. Haeberlen, Polymer9,51 (1968).58R.A. Assink, MacromoleculesAt, 1233(1978).59H. Wennerstrom, J. Magn. Reson. 11,219(1973).soj.S. Noggle and R.E. Schirmer, The Nuclear Overhauser Effect,

Academic Press, New York, 1971.61 J.R. Zimmerman and W.E. Brittin, J. Phys. Chem.61,1328(1957).62H.A. Resing, J. Chem. Phys.43,669(1965).63G.P. Jones, Phys. Rev. 148,332 (1966).MJ.F. Jacquinot and M. Goldman, Phys. Rev. 68,1944(1973).65B.A. Cornell and J.M. Pope, J. Magn. Reson. 16,172 (1974).66 J. Koppelman, Koll.-Z. Z Polym.216/217,6 (1967).67 K. Schmauder and R. Kimmich, unpublished results.6 8N. Bloembergen, S. Shapiro, P.S. Pershan, and J.O. Artman,

Phys. flei/114,445 (1959).69 P.S. Pershan, Phys. RevAM, 109(1960).70D.E. Woessner and H.S. Gutowsky, J. Chem. Phys. 29,804 (1958).71G. Voigtand R. Kimmich, J. Magn. Reson.24,149(1976).72A.G. Redfield, Phys. Rev. 130,589(1963).73 R.E. Slusher and E.L. Hahn, Phys. Rev. 166,332 (1968).74D.T. Edmonds and P.A. Speight, J. Magn. Reson. 6,265 (1972).75 D. Jerome and G. Galleron, J. Phys. Chem. Solids2i, 1557 (1963).76G.E. Pake and T.L. Estle, The Physical Principles of Electron

Paramagnetic Resonance, W.A. Benjamin, Reading, Mass., 1973.77F.A. Rushworth and D.P. Tunstall, Nuclear Magnetic Resonance,

Gordon and Breach, New York, 1973.78 G. Voigt and R. Kimmich, Progr. Colloid & Polym. Sci. 66, 273

(1979).79 M. Packard and R. Varian, Phys. Rev. 93,941 (1954).so A. Bloom and D. Mansir, Phys. Rev. 93,941 (1954).81R.V. Pound, Phys. Rev. 81,156(1951).82N.F. Ramsey and R.V. Pound, Phys. Rev.Bt, 278(1951).83R.T.Schumacher, Phys. Rev. 112,837(1958).84 B.C. Johnson and W.I. Goldburg, Phys. Rev. 145,380 (1966).85Z. Florkowski, I.W. Hennel, and B. Blicharska, Nucleonical*, 563

(1969).86R. Hausser, H. Kolb, and G. Siegle, Z angew. Phys.22,375 (1967).87G.P. Jones, J.T. Daycock, and T.T. Roberts, J. Sci. Instrum. (J.

Phys. E.) 2,630 (1969).88 A. L. Bloom, J. Chem. P/jys.25,793(1956).89H. Sprinz, Ann. Physik(Leipzig)20,168(1967).""A.G. Redfield, W. Fite, and H. E. Bleich, Rev. Sci. Instrum. 39, 710

(1968).9< K. Hallenga and S.H. Koenig, Biochemistry^, 4255 (1976).92 R. Kimmich and F. Noack, Z angew. Phys. 29,248 (1970).93E.v. Goldammer and W. Kreysch, Ber. Bunsenges. Phys. Chem.

82,463(1978).

94E. Dally and W. Muller-Warmuth, Ber. Bunsenges. Phys. Chem.82,792(1978).

95 M. S tohrer and F. Noack , J. Chem. Phys. 67 ,3729 (1977).96 N. Dav idson and R. Go ld , Biochim. Biophys. Acta26,370 (1957).97 O.K. Daszkiewicz, J .W. Hennel , B. Lubas, and T.W. Szczepkowsk i ,

/Vafc/ re(London)200,1006(1963).9 8 S . H . Koenig and W.E. Schi l l inger, J. Biol. Chem. 244, 3283 and

6520(1969).99B. Blicharska, Z. Florkowski, J.W. Hennel, G. Held, and F. Noack,

Biochim. Biophys. Acta207,381 (1970).100 R. Kimmich and F. Noack, Z Naturforsch. Teil A 25,299 and 1680

(1970).101R. Kimmich, and R. Noack, Ber. Bunsenges. Phys. Chem.75, 269

(1971).102R. Kimmich, Z Naturforsch. Teil B26,1168(1971).103S.H. Koenig, R.D. Brown, and C.F. Brewer, Proc. Natl. Acad. Sci.

l/&470,475(1973).104T.R. Lindstrom and S.H. Koenig, J. Magn. Reson. 15,344 (1974).105S.H. Koenig, K. Hallenga, and M. Shporer, Proc. Natl. Acad. Sci.

USA72,2667 (1975).106T.R. Lindstrom, S.H. Koenig, T. Boussios, and J.F. Bertles,

Biophys. J. 16,679 (1976).107L.Grosch, and F. Noack, Biochim. Biophys. Acta4S3 218(1976).108S.H. Koenig, R.G. Bryant, K. Hallenga and G.S. Jacob,

Biochemistry17,4348 (1978).109G. Jones, and J. Fornwalt, J. Chem. Phys. 4,30 (1936).110B.D. Hilton and R.G. Bryant, J. Magn. Reson.21,105(1976).111F. Gaggelli, E. Tiezzi, and G. Valensin, Chem. Phys. Lett. 63, 155

(1979).1121.D. Campbell and R. Freeman, J. Magn. Reson. 11,143(1973).113 H.T. Edzesand E.T. Samulski, /Vafc/re(London)265,521 (1977).114 A. Kalk and H.J.C. Berendsen, J. Magn. Reson. 24,346 (1976).115B.D.Sykes,W.E. Hull, and G.H.Snyder, Biophys. J.21,137(1978).116J.D. Stoesz, A.G. Redfield, and D. Malinowski, FEBS Letters 91,

320(1978).117H.T. Edzesand E.T. Samulski, J. Magn. Reson.31,207(1978).118M. Eisenstadt and M.E. Fabry, J. Magn. Reson. 29,591 (1978).119H.E. Bleich, K.R.K. Easwaran, and J.A. Glasel, J. Magn. Reson. 31,

517(1978).120M. Llinas, M.P. Klein, and K. Wuthrich, Biophys. J. 24,849 (1978).121F. Winter, Diplomarbeit, UniversitSt Ulm, in preparation.1 2 2J. Eisinger, R.G. Shulman, and B.M. Szymanski, J. Chem. Phys.36,

1721(1962).123 J. Eisinger, F. Fawaz-Estrup, and R.G. Shulman, J. Chem. Phys.

42,43(1965).124A.R. Peacocke, R.E. Richards, and B. Sheard, Mol. Phys. 16,177

(1969).125A.S. Mildvan and M. Conn, Adv. Magn. Reson. 33,1 (1970).126T.L. James, J. Reuben, and M. Cohn, J. Biol. Chem. 248, 6443

(1973).127 R.A. Dwek, Advan. Mol. Relaxation Processes*, 1 (1972).128J.M. Young, K.J. Schray, and A.S. Mildvan, J. Biol. Chem. 250,

9021(1975).129D.R. Burton, R.A. Dwek, S. Forsen, and G. Karlstrom,

Biochemistry^, 250 (1977).130 R.S. Levy and J.J. Villafranca, Biochemistry^, 3301 (1977).131 M.E. Fabry, J. Biol. Chem. 253,3568 (1978).132 B. Boettcher and M. Martinez-Carrion, J. Biol. Chem. 253, 4642

(1978).133T.R. Krugh, in Spin Labeling: Theory and Applications, L.J. Ber-

liner, Ed., Academic Press, New York, 1976, p. 339.134 B.P. Gaber, W.E. Schillinger, S.H. Koenig, and P. Aisen, J. Biol.

Chem. 245,4251 (1970).135M.E. Fabry, S.H. Koenig, and W.E. Schillinger, J. Biol. Chem. 245,

4256(1970).

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

Vol. 1,No. 4 217

Page 24: Field Cycling in NMR Relaxation Spectroscopy: Applications ...

«6S.H. Koenig and R.D. Brown. Ann. N.Y.Acad. Sci.222,752(1973).137 S.H. Koenig, R.D. Brown, and C.F. Brewer, Proc. Natl. Acad. Sci.

US/470,475 (1973).138R.D. Brown, C.F. Brewer, and S.H. Koenig, Biochemistry"16,3883

(1977).139D. Chapman, Ed., Biological Membranes, Academic Press, Lon-

don and New York, 1968.""H.Trauble, J. Membr. 6/O/.4,193(1971).141 J.T. Daycock, A. Darke, and D. Chapman, Chem. Phys. LipidsB,

205(1971).142J.H. Davis, K.R. Jeffrey, and M. Bloom, J. Magn. Reson. 29, 191

(1978).143 Y.K. Levine, N.I.M. Birdsall, A.G. Lee, and I.C. Metcalfe,

Biochemistry^, 1416(1972).144 N.J. Salsbury, D. Chapman, and G.P. Jones, Trans. Faraday Soc.

66,1554(1970).145 R. Kimmich and G. Voigt, Chem. Phys. Lett. 62,181 (1979).146 A. PetersTThesis, Universitat Ulm, 1979.147P.T. Beall, C.F. Hazlewood, and P.N. Rao, Sc/ence 192,904(1976).148 R. Damadian, Science1T\, 1151 (1971).149 D.P. Hollis, Bull. Magn. Reson. 1,27 (1979).150 L.E. Barroilhet and P.R. Moran, Med. Phys. 3,410 (1976).151H. Grasdalen, L.E.G. Eriksson, A. Ehrenberg, and D. B3ckstr6m,

Biochim. Biophys. Acta 541,521 (1978).152C.F. Hazlewood, D.C. Chang, B.L. Nichols, and D.E. Woessner,

Biophys. J. 14,583 (1974).153M.M. Civan and M. Shporer, Biophys. J. 15,299 (1975).154 J.G. Diegel and M.M. Pintar, Biophys. J. 15,855 (1975).155A. Zipp, I.D. Kuntz, and T.L. James, J. Magn. Reson. 24, 411

(1976).156K.R.' Foster, H.A. Resing, and A.N. Garroway, Science 194, 324

(1976)..157 D.C. Chang and D.E. Woessner, Science 198,1180 (1977).158M.M. Civan, A.M. Achlama, and M. Shporer, Biophys. J. 21, 127

(1978). '•IK> G. Held, F. Noack, V. Pollak, and B. Melton, Z. Naturforsch. Teil C

28,59(1973).160 B.M. Fung, Biophys. J. 18,235 (1977).161P. Pincus, Solid State Commun. 7,415 (1969).162J.W. Doane, C.E. Tarr, and M.A. Nickerson, Phys. Rev. Lett. 33,

620(1974).163T.E. Faber, Proc. R. Soc. London Ser. A 353, 247, 261, and 277

(1977).164S.D. Goren, C. Korn, and S.B. Marks, Phys. Rev. Lett 34, 1212

(1975).165R. Blinc, M. Vilfan, and V. Rutar, Solid State Commun. 17, 171

(1975).166B.M. Fung, C.G. Wade, and R.D. Orwoll, J. Chem. Phys. 64, 148

(1976).167B.M. Fung andT.A. Martin, J. Chem. Phys.61,1698(1974).168 w . Wolfel, F. Noack, and M. Stohrer, Z. Naturforsch. Teil A 30,437

(1975).<89V. Graf, F. Noack, and M. Stohrer, Z Naturforsch. Teil A 32, 61

(1977).170 w. Wolfel and F. Noack, in Magnetic Resonance and Related

Phenomena, E. Kundla, E. Lippmaa, and T. Saluvere, Eds., Springer-Verlag, Berlin, Heidelberg, and New York, 1979, p. 4"55.

171 M.F. Brown, J. Magn. Reson.35,203 (1979).172R.G. Brussau and H. Sillescu, Ber. Bunsenges. Phys. Chem.76,31

(1972).173 J.J. v.d. Klink, J. Schriever, and J.C. Leyte, Ber. Bunsenges. Phys.

Chem. 78,369 (1974).1 7 4J. Schriever, S.W.T. Westra, and J.C. Leyte, Ber. Bunsenges.

. Phys. Chem. 81,287 (1977).1 7 5J. Schriever, J.J. Bunk, and J.C. Leyte, J. Magn. Reson. 27, 45

(1977).176 Y. Inoue and T. Konno, Polym. J. 8,457 (1976).177 W, Gronski and N. Murayama, Makromol. Chem. 180,277(1979).178 H-C. Broecker and J. Klahn, Makromol. Chem. 180,551 (1979).'79R. Ullman, J. Chem. Phys.43,3161 (1965).180 R. Ullman, J. Chem. Phys.AA, 1558(1966).181W.L.F. Golz and H.G. Zachmann, Makromol. Chem. 176, 2721

(1975).182 R. Lenk and J.P. Cohen-Addad, Solid State Commun. 8, 1864

(1970).183G. Preissing and F. Noack, Progr. Colloid & Polym. Sci. 57, 216

(1975).184 H. Koch and R. Kimmich, Polymer20,132 (1979).185 R. Kimmich and H. Koch, Colloid & Polym. Sci., in press.186R. Kimmich, Polymer^, 233 (1977).187 R. Kimmich and K. Schmauder, PolymerW, 239 (1977).188 H. Koch, R. Bachus, and R. Kimmich, unpublished results.189 B. Valeur, J.P. Jarry, F. Geny, and L. Monnerie, J. Polym. Sci.,

Polym. Phys. ed., 13,667 (1975).190 A.A. Jones and W.H. Stockmayer, J. Polym. Sci., Polym. Phys. ed.,

15,847(1977).191 N.J. Trappeniers, C.J. Gerritsma, and P.H. Oosting, Physica 30,

997(1964).192 U. Haeberlen, Koll.-Z. Z. Polym. 225,15 (1968).193T.M. Connor, D.J. Blears, and G. Allen, Trans. Faraday Soc. 61,

1097(1965).194 B. Crist and A. Peterlin, J. Macromol. Sci. Phys. 84, 791 (1970).195T.J. Rowland and L.C. Labun, Macromoleculesti, 466 (1978).196G. Voigt and R. Kimmich, Chem. Phys. Lett. 56,275 (1978).197V.J. McBrierty, D.W. McCall, D.C. Douglass, and D.R. Falcone, J.

Chem. Phys. 52,512 (1970).198 W.P . Sl ichter , J. Chem. Ed AT, 193(1970).199 Q voigt and R. Kimmich, unpublished results.200 K. Meise, W. Muller-Warmuth, and H.W. Nientiedt, Ber. Bun-

senges. Phys. Chem. 80,584 (1976).201G. Maier and H.C. Wolf, Z Naturforsch. Teil A 23,1068 (1968).202 S.H. Koenig and M. Epstein, J. Chem. Phys. 63,2279 (1975).203 H. Pfeifer, Z Naturforsch. Teil A17,279 (1962).204 R. Hausser and F. Noack, Z Naturforsch. Teil A 20,1668 (1965).205 E. v. Goldammer, W. Kreysch, and H. Wenzel, J. Solution. Chem. 7,

197(1978).206 E. Dally and W. Mu l l e r -Warmu th , Ber. Bunsenges. Phys. Chem. 82,

792(1978).2 0 7H. Kolb and H.C. Wolf, J. Magn. Reson.7,374 (1972).2°8 H. Kolb and H.C. Wolf, Z Naturforsch. Teil A 27,51 (1972).2 0 9J. Winter, Magnetic Resonance in Metals, Oxford University

Press, Oxford, 1971, p. 141.210 A.Z. Genack and A.G. Redfield, Phys. Rev. 612,78 (1975).211A.Z. Genack, Phys. Rev. 613,68(1976).

Duplication of Bulletin of Magnetic Resonance, in whole or in part by any means for any purpose is illegal.

218 Bulletin of Magnetic Resonance

1• l !