intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e.,...

12
Intermediate motions as studied by solid-state separated local field NMR experiments Eduardo Ribeiro deAzevedo, 1,a Kay Saalwachter, 2 Ovidiu Pascui, 2 André A. de Souza, 1 Tito J. Bonagamba, 1 and Detlef Reichert 2 1 Instituto de Física de São Carlos, Universidade de São Paulo, Caixa Postal 369, CEP, 13560-970 São Carlos, São Paulo, Brazil 2 Institut für Physik, Martin-Luther-Universität Halle-Wittenberg, Friedemann-Bach-Platz 6, D-06108 Halle, Germany Received 28 August 2007; accepted 13 December 2007; published online 13 March 2008 In this report, the application of a class of separated local field NMR experiments named dipolar chemical shift correlation DIPSHIFT for probing motions in the intermediate regime is discussed. Simple analytical procedures based on the Anderson–Weiss AW approximation are presented. In order to establish limits of validity of the AW based formulas, a comparison with spin dynamics simulations based on the solution of the stochastic Liouville–von-Neumann equation is presented. It is shown that at short evolution times less than 30% of the rotor period, the AW based formulas are suitable for fitting the DIPSHIFT curves and extracting kinetic parameters even in the case of jumplike motions. However, full spin dynamics simulations provide a more reliable treatment and extend the frequency range of the molecular motions accessible by DIPSHIFT experiments. As an experimental test, molecular jumps of imidazol methyl sulfonate and trimethylsulfoxonium iodide, as well as the side-chain motions in the photoluminescent polymer poly2-methoxy-5- 2-ethylhexyloxy-1,4-phenylenevinylene, were characterized. Possible extensions are also discussed. © 2008 American Institute of Physics. DOI: 10.1063/1.2831798 INTRODUCTION Due to its strong dependence on the mutual distance be- tween the interacting spins, the strength of the heteronuclear dipolar coupling is a valuable source for structural informa- tion. In order to resolve different chemical moieties in the NMR spectra, these distance measurements generally require magic-angle spinning MAS. The magnitude of the dipolar coupling is, however, not only a valuable source to gain structural information but also for extracting dynamic param- eters. The ratio between the experimentally determined re- duced dipolar coupling and the rigid-lattice value, the order parameter, indicates fast molecular processes that average and, consequently, reduce the dipolar interaction. 1 Besides the simple dipolar line-shape 2 and relaxation experiments, some more elaborate approaches deserve special attention, namely, cross-polarization CP dynamics 3 and separated lo- cal field 4,5 SLF experiments. The first relies on the hetero- nuclear dipolar coupling between abundant such as 1 H and rare spins such as 13 C in CP experiments. The increase of the rare-spin magnetization occurs with the time constant T CH that is—similar to the relaxation time T 1 —sensitive to dynamic processes in the kilohertz range and correlates to important mechanical parameters such as dynamic moduli. The measurement of T CH became rather popular in dynamic investigations of organic solids. 6 However, care must be taken in the interpretation of the data. A more straightfor- ward experiment is the Lee–Goldburg LG-CP, 7 in which the homonuclear 1 H dipolar interaction is suppressed and the buildup curve of the rare-spin magnetization is governed by the 1 H– 13 C interaction. SLF experiments provide a correla- tion between the chemical shifts and dipolar couplings. A popular version of SLF experiments is the wide-line separa- tion, which was first suggested as a dynamic filter to separate mobile and rigid components in 13 C spectra 8 and was later turned into an easy-to-implement two-dimensional method. 9 However, these experiments suffer from the simultaneous ac- tion of heteronuclear 13 C– 1 H and homonuclear 1 H– 1 H di- polar interactions, the latter masking the clear information provided by the former. An SLF variant that deserves par- ticular attention is the efficient dipolar chemical shift corre- lation DIPSHIFT sequence. 5,10,11 All these experiments de- termine the dipolar coupling strength that might be scaled down by fast molecular dynamics occurring with correlation times C within the range from 10 -5 to 10 -4 s. Information about the amplitude of the molecular process can be ex- tracted from the order parameter reduction and most of the experiments mentioned above might be used for this pur- pose. The DIPSHIFT pulse sequence was identified to be a particularly robust experiment and is going to be used throughout this paper. Due to this reason, both the theory and simulations presented below are dedicated to this experi- ment. The most important issue for studying molecular dy- namic processes is to characterize both the motion ampli- tudes and time constants over a wide dynamic range, involv- ing the application of several dynamic NMR methods. As discussed above, the usual applications of the SLF experi- ments are performed in the fast-motion regime, where they a Author to whom correspondence should be addressed. Electronic mail: [email protected]. THE JOURNAL OF CHEMICAL PHYSICS 128, 104505 2008 0021-9606/2008/12810/104505/12/$23.00 © 2008 American Institute of Physics 128, 104505-1 Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Upload: others

Post on 20-Aug-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e., i it must permit the appli-cation of MAS and 1H decoupling during the acquisition

Intermediate motions as studied by solid-state separated local fieldNMR experiments

Eduardo Ribeiro deAzevedo,1,a� Kay Saalwachter,2 Ovidiu Pascui,2 André A. de Souza,1

Tito J. Bonagamba,1 and Detlef Reichert21Instituto de Física de São Carlos, Universidade de São Paulo, Caixa Postal 369, CEP,13560-970 São Carlos, São Paulo, Brazil2Institut für Physik, Martin-Luther-Universität Halle-Wittenberg, Friedemann-Bach-Platz 6,D-06108 Halle, Germany

�Received 28 August 2007; accepted 13 December 2007; published online 13 March 2008�

In this report, the application of a class of separated local field NMR experiments named dipolarchemical shift correlation �DIPSHIFT� for probing motions in the intermediate regime is discussed.Simple analytical procedures based on the Anderson–Weiss �AW� approximation are presented. Inorder to establish limits of validity of the AW based formulas, a comparison with spin dynamicssimulations based on the solution of the stochastic Liouville–von-Neumann equation is presented. Itis shown that at short evolution times �less than 30% of the rotor period�, the AW based formulasare suitable for fitting the DIPSHIFT curves and extracting kinetic parameters even in the case ofjumplike motions. However, full spin dynamics simulations provide a more reliable treatment andextend the frequency range of the molecular motions accessible by DIPSHIFT experiments. As anexperimental test, molecular jumps of imidazol methyl sulfonate and trimethylsulfoxonium iodide,as well as the side-chain motions in the photoluminescent polymer poly�2-methoxy-5-�2�-ethylhexyloxy�-1,4-phenylenevinylene�, were characterized. Possible extensions are alsodiscussed. © 2008 American Institute of Physics. �DOI: 10.1063/1.2831798�

INTRODUCTION

Due to its strong dependence on the mutual distance be-tween the interacting spins, the strength of the heteronucleardipolar coupling is a valuable source for structural informa-tion. In order to resolve different chemical moieties in theNMR spectra, these distance measurements generally requiremagic-angle spinning �MAS�. The magnitude of the dipolarcoupling is, however, not only a valuable source to gainstructural information but also for extracting dynamic param-eters. The ratio between the experimentally determined re-duced dipolar coupling and the rigid-lattice value, the orderparameter, indicates fast molecular processes that averageand, consequently, reduce the dipolar interaction.1 Besidesthe simple dipolar line-shape2 and relaxation experiments,some more elaborate approaches deserve special attention,namely, cross-polarization �CP� dynamics3 and separated lo-cal field4,5 �SLF� experiments. The first relies on the hetero-nuclear dipolar coupling between abundant �such as 1H� andrare spins �such as 13C� in CP experiments. The increase ofthe rare-spin magnetization occurs with the time constantTCH that is—similar to the relaxation time T1�—sensitive todynamic processes in the kilohertz range and correlates toimportant mechanical parameters such as dynamic moduli.The measurement of TCH became rather popular in dynamicinvestigations of organic solids.6 However, care must betaken in the interpretation of the data. A more straightfor-ward experiment is the Lee–Goldburg �LG�-CP,7 in which

the homonuclear 1H dipolar interaction is suppressed and thebuildup curve of the rare-spin magnetization is governed bythe 1H– 13C interaction. SLF experiments provide a correla-tion between the chemical shifts and dipolar couplings. Apopular version of SLF experiments is the wide-line separa-tion, which was first suggested as a dynamic filter to separatemobile and rigid components in 13C spectra8 and was laterturned into an easy-to-implement two-dimensional method.9

However, these experiments suffer from the simultaneous ac-tion of heteronuclear 13C– 1H and homonuclear 1H– 1H di-polar interactions, the latter masking the clear informationprovided by the former. An SLF variant that deserves par-ticular attention is the efficient dipolar chemical shift corre-lation �DIPSHIFT� sequence.5,10,11 All these experiments de-termine the dipolar coupling strength that might be scaleddown by fast molecular dynamics occurring with correlationtimes �C within the range from 10−5 to 10−4 s. Informationabout the amplitude of the molecular process can be ex-tracted from the order parameter reduction and most of theexperiments mentioned above might be used for this pur-pose. The DIPSHIFT pulse sequence was identified to be aparticularly robust experiment and is going to be usedthroughout this paper. Due to this reason, both the theory andsimulations presented below are dedicated to this experi-ment.

The most important issue for studying molecular dy-namic processes is to characterize both the motion ampli-tudes and time constants over a wide dynamic range, involv-ing the application of several dynamic NMR methods. Asdiscussed above, the usual applications of the SLF experi-ments are performed in the fast-motion regime, where they

a�Author to whom correspondence should be addressed. Electronic mail:[email protected].

THE JOURNAL OF CHEMICAL PHYSICS 128, 104505 �2008�

0021-9606/2008/128�10�/104505/12/$23.00 © 2008 American Institute of Physics128, 104505-1

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 2: Intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e., i it must permit the appli-cation of MAS and 1H decoupling during the acquisition

are only sensitive to the motion amplitudes. By slowingdown the dynamic process by decreasing the temperature,one can reach the so-called intermediate-motion regime, atwhich the time constant of the dynamic process is similar tothe inverse coupling constant of the anisotropic interaction.12

This scenario, which is mathematically very similar to thetreatment of exchange NMR experiments,13 is well known inliquid-state NMR where the exchange of different isotropicfrequencies leads to a loss of phase coherence and to abroadening of the spectra. The same approach is applicablefor static solid-state spectra, where the exchange occurs be-tween different anisotropic frequencies, i.e., between nucleibelonging to molecules with different orientations with re-spect to the external magnetic field. The approach is welldeveloped and was applied frequently to 2H spectra in whichthe quadrupolar interaction is the dominating source of linebroadening, making the calculation of the effect of moleculardynamic processes on the line shape relatively easy andstraightforward.14,15 With MAS, however, a more elaboratedtheory is necessary to deal with the additional time depen-dence of the resonance frequencies, imposed by the MASrotation. Qualitatively, the incoherent exchange processes in-terfere with the coherent time dependence of the MAS andspoil the periodicity of the time-domain signal, leading to acharacteristic line broadening.16 Successful approaches tocalculate spectra include the application of the Anderson–Weiss theory,17–20 Floquet theory,15,21 extended exchangetheory,22,23 or stepwise numerical propagation in the frame-work of partial analytical solutions or of explicit density ma-trix simulations.24,25 In addition, in most of the cases fornatural isotopic abundance nuclei in organic solids, dipolardecoupling has to be applied in concert with MAS to achieveappreciable spectral resolution. The 1H irradiation adds an-other coherent perturbation that interferes with the incoher-ent molecular exchange process and results in another sourceof line broadening. This effect was realized26 early and al-ready put to use qualitatively.27 However, it is difficult toquantitatively simulate the spectra because it involves therelatively easy-to-handle chemical shift anisotropy �CSA� in-teraction as well as the more complicated hetero- and homo-nuclear dipolar interactions. Also, the actual line broadeningspoils the spectral resolution, which is highly undesirable butpartially unavoidable.

This prompts for a dynamic experiment that works in theintermediate-motional regime under the condition of naturalabundance in organic solids, i.e., �i� it must permit the appli-cation of MAS and 1H decoupling during the acquisitionperiod, �ii� should use a single and well-defined anisotropicinteraction for probing the molecular dynamics, and �iii� forissues of spectral resolution, should be as little sensitive todynamic line broadening as possible. In this paper we sug-gest the application of SLF experiments that correlate the13C chemical shift resonances in a normal CP-MAS spec-trum with the heteronuclear dipolar 1H– 13C couplings. Sincethe 1H– 1H coupling is efficiently suppressed by homo-nuclear decoupling sequences such as the LG28 or its ad-vanced variants frequency switched Lee–Goldburg29 andphase modulated Lee–Goldburg,30 the 1H– 13C interaction ispredominant between a given 13C and the directly bonded 1H

for aliphatic carbons. This interaction is axially symmetricand oriented along the C–H bond, which makes it a veryconvenient probe for the dynamic process. The poor spectralresolution is still an issue but can be partially worked aroundby the application of higher MAS rates that partially quenchthe line broadening effect. However, this also attenuates theheteronuclear dipolar coupling and makes the experimentless sensitive to the molecular dynamic processes. In turn,this can be partially overcome by the application of a DIP-SHIFT variant that employs the recoupling of the 13C– 1Hdipolar interaction, i.e., measuring a multiple of the actualinteraction strengths.11 Another disadvantage of the DIP-SHIFT experiments is found when cross-polarization is usedas the excitation scheme. The presence of internal motionsduring the contact time produces signal reduction. This prob-lem becomes more severe when distribution of correlationtimes is present. In this case only part of the moving seg-ments effectively contributes to the DIPSHIFT curves, de-creasing the accuracy of the parameters measured underthese circumstances. Alternatively, the intrinsic advantage ofthe SLF experiments is that only a specific interaction con-tributes to the signal and no extra experiment for data nor-malization is necessary.

In this report, we present the theory as well as experi-mental results for DIPSHIFT experiments applied in theintermediate-motion regime. For the extraction of kinetic pa-rameters, the comparison of the experiments with calculateddata is necessary. We present the numerical approach anddiscuss the influence of magnetic parameters of the sampleas well as dynamic parameters on the calculated data. Inaddition, the accuracy and limits of validity of a fitting for-mula derived based on the Anderson–Weiss approach18,19 isdiscussed. The approach is applied to molecular jumps ofimidazol methyl sulfonate and trimethylsulfoxonium iodide�TMSI�, as well as to the side-chain motion in the photolu-minescent polymer poly�2-methoxy-5-�2�-ethylhexyloxy�-1,4-phenylenevinylene� �MEH-PPV�. Finally, we addresspossible extensions.

METHOD AND MATERIALS

The pulse sequence for the DIPSHIFT experiment isshown in Fig. 1�a�. After the excitation procedure �CP, Directpolarization, etc.�, the magnetization evolves during a vari-able time t1 under the influence of chemical shift and hetero-nuclear dipolar interaction, while the 1H homonuclear cou-pling is suppressed by 1H LG irradiation implemented in aphase modulated fashion30 �PMLG�. The Hahn echo at t2

=0 removes the effect of chemical shift and leaves the1H– 13C dipolar coupling as the only effective interactionduring the evolution time t1. The signal detection under MASduring t2 is performed for different t1 values. Thet1-dependent modulation of the resulting line intensities inthe spectra obtained can be used to extract the value of the1H– 13C dipolar coupling. In the case of fast dynamic pro-cesses, a reduced coupling is measured and can be turnedinto an order parameter of motion that in turn can be relatedto jump angles if a certain motional geometry is known orproposed, see Fig. 1�b�. The models presented reflect the

104505-2 deAzevedo et al. J. Chem. Phys. 128, 104505 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 3: Intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e., i it must permit the appli-cation of MAS and 1H decoupling during the acquisition

most common anisotropic motions, two- �1�, three- �2�, andfour- �3� site jumps and rotation on a cone �4,5�. Note thatfor increasing reorientation angle �, the order parameter firstdecreases to small values �0.2 at �=90° and then increasesto 1 at �=180°. Such periodic behavior reflects the symmetryof the �3 cos2 �−1� angle dependence. However, this behav-ior is quite similar for all models, which makes the accessedmotion geometry information not very reliable.

It is important to emphasize that the extraction of param-eters related to the actual molecule, such as jump angles,requires a motional model that cannot always be fully justi-fied on the basis of limited experimental data. An extractedaveraged dipolar coupling and thus the experimental orderparameter of motion can be explained well by different jumpangles for different models, as demonstrated in Fig. 1�b�.Thus, results of these experiments are always, to a certaindegree, ambiguous.

The pulse sequence shown in Fig. 1�a� was applied un-changed for the experiments described in this paper. Theexperiments were performed on a VARIAN INOVA 400spectrometer, equipped with a VARIAN 7 mm VT-CPMASprobe �Jacobsen design�. Typical rf parameters are 70 kHz�3.6 �s � /2 proton pulse� during the LG irradiation �actuallyPMLG30 pulse sequence� and 60 kHz for heteronuclear dipo-lar decoupling.

Figure 2 shows the chemical structures of the materialsstudied in this paper. Imidazol methyl sulfonate is an ex-ample for a well-defined two-site jump process and was al-ready the subject of dynamic investigations,31 so that therelevant kinetic parameters are known and can be used toevaluate the method. This sample was provided by Sluchter.TMSI is a test substance in which molecules undergo a well-defined three-site jump motion and fast CH3 rotation thatproduces a preaveraged dipolar coupling. In contrast to theimidazol derivative, it permits carrying out experiments overthe full dynamic range of the method without reaching themelting point. MEV-PPV is a semiconducting polymer in

which the molecular dynamic processes directly affect itsphotoluminescent properties.32,33 In particular, it has beensuggested that the MEH-PPV side chains perform three-sitejumps with a distribution of correlation times. The MEH-PPV sample was obtained from Aldrich and used as receivedin the form of freestanding films cast from a concentratedchloroform solution.

THEORY

Spin dynamics simulations

The quantification of the DIPSHIFT data in theintermediate-motion regime can be performed using manyapproaches that describe the effects in this regime underMAS. This includes the Floquet theory,15,21 extended ex-change theory,22,23 approximate theories based on theAnderson–Weiss theory,18–20 and stepwise numericalpropagation24,25 using the solution of the stochasticLiouville–von-Neumann equation. Among these approaches,full spin dynamics simulations based on density matrixpropagation were chosen to be employed here for the methodvalidation. Specifically, a spin dynamics simulation programdescribed by Saalwachter and Fischbach in Ref. 25 was used.Since there is a comprehensive description of this approachin Ref. 25, the details of the simulation method are not goingto be described here but only the main features that relate tothe adaptation to our specific problem. Simply put, the simu-lation approach basically consists of a multistep calculationof the density matrix evolution under the effect of quantum-mechanical evolution under MAS and dynamic exchange. Aspecific feature of the program is that the quantum-mechanical �propagator-based� evolution of individual den-sity matrices related to the different sites and mixing stepsbetween the sites mediated by an exchange matrix are per-formed independently in subsequent steps. This avoids thecomputationally costly diagonalization of the full Liouville-space propagator that embodies both quantum mechanics andexchange steps. In the particular case of the DIPSHIFT pulse

FIG. 1. �a� DIPSHIFT pulse sequence. �b� Depiction of some selected mo-tional geometries, and �c� the behavior of the corresponding order param-eters as a function of the motional amplitude �opening angle�.

FIG. 2. Molecular structures of �a� imidazole methyl sulphonate,�b� trimethylsulfoxonium iodide �TMSI�, and �c� �2-methoxy-5-�2�-ethylhexyloxy�-1,4-phenylenevinylene� �MEH-PPV�.

104505-3 Intermediate motions in solid-state NMR experiments J. Chem. Phys. 128, 104505 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 4: Intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e., i it must permit the appli-cation of MAS and 1H decoupling during the acquisition

sequence, the evolution of CH �two spin system�, CH2 �threespin system�, or CH3 �four spin system� groups during t1 wassimulated considering off-resonance irradiation of the pro-tons according to the LG conditions,28 i.e., the frequencyoffset was �=1 /�2 so that the effective field is 54.7°away of the magnetic field direction. Thus, the scaling factorintroduced by the use of the LG or PLMG irradiation is takeninto account in the simulation. Because of the rotor-synchronized � pulses, the chemical shifts are refocused inthe pulse sequence and are not considered in the simulations.Another detail concerns the simulation of CH3 groups, whichwas performed considering an average orientation of the pro-tons among the three symmetric positions around the C3

symmetry axis. This is equivalent to a fast rotation of the CHbonds around the C3 axis.

To demonstrate the approach, we turn to the particularsimple case of a two-site jump, as performed by the imida-zole molecule, Fig. 2�a�. Figure 3 shows theoretical DIP-SHIFT dephasing curves, calculated with the parameters thatdescribe the dynamics of basal carbons in imidazole methylsulfonate31,34 �CH dipolar coupling DCH=23 kHz and reori-entation angle=64°�, for different molecular jump rates. Thecurve for 1101 Hz is almost in the rigid limit �slow-motion� case and yields the full 1H– 13C dipolar coupling�scaled by the LG irradiation�, while the curve for 1106 Hz corresponds to the fast-motion limit with a dipolarcoupling averaged by the two-site jump of this molecule.The effective dipolar couplings in these cases are roughlyproportional to the depth of the dephasing curve, i.e.,I�t1=0.5TROT�. Note that in these cases, the intensity at t1

=TROT evolves back to the value for t1=0, indicating that theperiodicity of the Hamiltonian due to the MAS rotation ispreserved. For all jump rates between these limiting cases,the dephasing curve intensity at t1=TROT does not reach its

initial value due to the destructive interference of the mo-lecular dynamics, i.e., the reorientation of the dipolar vectorunder MAS rotation. We can consider this as an increasingdynamic contribution to the overall T2 value or, in terms ofthe spectral picture, as a dynamic broadening of a fictitious1H– 13C dipolar spectrum. In particular, it is worth mention-ing that for jump rates smaller than 1104 Hz �Fig. 3�a�� themain effects observed are �i� a shortening of the apparent T2

times and the constancy of the apparent dipolar couplings,and for jump rates larger than 1104 Hz �Fig. 3�a��, themain effect observed is �ii� an increase of the apparent T2 inconcert with a decrease in the apparent dipolar couplings�Fig. 4�a��. The details of these changes depend on the ge-ometry of the process �jump angles, number of sites, etc.�, aswill be demonstrated below. In particular, this is a crucialissue when the dynamic process exhibits a distribution ofjump rates/correlation times, as discussed below for theMEH-PPV case.

Analytical approaches based on the Anderson–Weissapproximation

The simulation procedure described above is very reli-able and can be used to demonstrate the feasibility of usingDIPSHIFT to study motions in the intermediate regime limit.However, there are some analytical treatments that are usedto describe the evolution of nuclear spins under the simulta-neous action of internal interactions �local fields�, intermedi-ate regime motions, and MAS. Because such approaches areeasier to implement and can be used in a straightforwardanalysis, it is worthwhile to test the specific conditions underwhich they can be applied. Here we are going to discuss theanalytical treatments based on the so-called Anderson–Weiss�AW� approximation, which has been successfully applied todescribe the evolution of static and MAS spectra under theeffect of diffusive motions.19,20 The first assumption of theAW approximation is that the distribution of local fields inthe sample �due to the internal NMR interactions� corre-sponds to a stationary Gaussian frequency distribution. Fur-ther, fluctuations of these local fields are assumed to be fullyrandom, and the stochastic loss of correlation is described bya monoexponential correlation function gtherm�t�=exp�−t /�C�, characterized by a single correlation time �C.Note that spin diffusion in multispin system, such as 1H nu-clei in polycrystalline or amorphous samples, usually pro-duces Gaussian frequency distributions �Van Vleck theory35�and monoexponential correlation functions. This makes theAW approach particularly successful when employed for de-scribing the dynamic evolution of 1H nuclei in organic pow-ders, where the multispin local fields lead to a “Gaussianiza-tion” and therefore to correlation loss by spin diffusion.36

Besides the stochastic loss of correlations assumed in theAW treatment, it is also necessary to introduce the coherentmodulation due to the MAS. This was already done in sev-eral articles16,18–20 considering local fields produced by eitherchemical-shift anisotropy or dipolar couplings. A particularlysuccessful approach was published first by Clough andGray,18 further developed by Fenzke et al.19 and extended toconsider diffusive anisotropic motion by Hirschinger.20 In

FIG. 3. Spin dynamics simulations of DIPSHIFT patterns at various jumprates k. The considered geometry of the motion was a two-site jump�k= �2�c�−1� with a reorientation angle of 64°, as found for the basal carbonsof imidazole methyl sulfonate.

104505-4 deAzevedo et al. J. Chem. Phys. 128, 104505 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 5: Intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e., i it must permit the appli-cation of MAS and 1H decoupling during the acquisition

the latter case, i.e., diffusive anisotropic motions with orderparameter S2= �M2�av /M2

rigid �final plateau value of the corre-lation function�, the NMR signal �free induction decay� ofdipolar coupled spins under the influence of MAS and mo-lecular dynamics, was found to be

S�t� = exp�− M2rigid�S2� 2

3 f�R,0,t� + 13 f�2R,0,t��

+ �1 − S2�� 23 f�R,1/�C,t� + 1

3 f�2R,1/�C,t��� ,

f�,1/�C,t� =t�C

1 + ��C�2 −�C

2 �1 − ��C�2��1 + ��C�2�2

1 − exp�−t

�C�cos�t�

−2�C

3

�1 + ��C�2�2 exp�−t

�C�sin�t� , �1�

where R is the angular MAS frequency, �C is the motioncorrelation time, and M2= �D

2 � is the second moment of theGaussian local field distribution. For the case where LG ir-radiation is used and a single CH pair is considered, M2

= �D2 �= 9

20�2�Dapp�2 with Dapp being the apparent dipolarcoupling constant, i.e., the dipolar coupling constant multi-plied by the LG scaling factor of 0.577. The use of the aboveformula allows us to calculate the MAS signal under theeffect of molecular motions. Thus, if one considers the localfield as being produced by the scaled C–H dipolar interac-tion, the points in S�t� covering the range from 0 to TROT

describes the DIPSHIFT curves. However, because bothmodels are based on the AW approximations and were re-stricted to small-step diffusive processes, it is expected thatthey might fail if these conditions are not fulfilled. Note thatthe Gaussian frequency distribution might be approximately

FIG. 4. �Color� �a� Comparison be-tween simulations of DIPSHIFT pat-terns in the rigid- and fast-motion lim-its at two spinning frequenciesobtained by different approaches: spindynamics simulations �points�, Eq. �1��solid lines�, and MAS equations�dashed lines�. �b� Spin dynamicssimulation for CH segments executingmotions with distinct numbers of sites�2 /3 /8 /18� on a cone of 64° aperture.�c� Comparison between the spin dy-namics and Eq. �1� based simulationsof DIPSHIFT patterns for �R=8 kHzand “3 /8 /18” site jumps and variousmotional rates �left�.

104505-5 Intermediate motions in solid-state NMR experiments J. Chem. Phys. 128, 104505 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 6: Intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e., i it must permit the appli-cation of MAS and 1H decoupling during the acquisition

provided by the “powder average,” but the limit of validityof that must be verified. Concerning the restriction to diffu-sive motions, it would also be interesting to check whetherthis produces only a constant difference in the correlationtimes or if it really changes the shape of the DIPSHIFTcurves. To verify this, we compare the DIPSHIFT curvesobtained using the spin dynamic simulations and Eq. �1�. Theresults will be described in the following section.

RESULTS AND DISCUSSIONS

Comparison between analytical approachesand spin dynamics simulations

In this section, DIPSHIFT curves obtained by spin dy-namics simulations and analytical approaches based on AWapproximations are compared. Since the simulation of themolecular motions in the spin dynamics program is per-formed by assuming a finite number of jumps between non-magnetically equivalent molecular sites, the motion geom-etry of imidazole methyl sulfonate, two-site jumps on a coneof 64° aperture, will be considered, see Fig. 2. Figure 4�a�shows comparisons among the DIPSHIFT curves obtainedby the spin dynamics simulation and Eq. �1� in the rigid andfast limits. In the rigid limit, a simulation using the standardMAS formulas37 is also appended. The curves obtained byspin dynamic simulations and MAS equations were obtainedconsidering a CH coupling strength DCH=22.7 kHz �C–Hdistance of 1.1 Å� which leads to Dapp= �0.577�DCH

=13.1 kHz and M2=3109 �rad /s�2. The calculation usingEq. �1� was performed using a rigid limit second moment ofM2

rigid=2.9109 �rad /s�2, which was chosen to fit the lowervalue of the spin dynamics simulated DIPSHIFT curve in therigid limit. The order parameter �S=0.55� was adjusted toproduce the same curve as the spin dynamics simulation inthe fast limit. At a spinning frequency of 4 kHz, a goodagreement is observed between the rigid limit patterns ob-tained by the spin dynamics program and MAS equation, buta considerable deviation is observed for the pattern obtainedfrom Eq. �1�. In principle, one could think that it would bedue to the value of M2

rigid, but even if this parameter is setfreely, the shape of the curves remains quite different. How-ever, the discrepancy is much smaller in the patterns ob-tained at 8 kHz spinning frequency. This behavior can beunderstood by considering the rule of the Gaussian fre-quency distribution required in the AW approximation. Thiscondition is actually not fulfilled in the case of isolated C–Hpairs spinning at 4 kHz because the shape of the stationaryfrequency distribution �MAS spectrum� still shows a Pake-like pattern, and therefore does not have any similarity witha Gaussian pattern, as considered in the calculations with Eq.�1�. When the spinning frequency is increased to 8 kHz, thereduction of the sidebands makes the Gaussian approxima-tion more realistic, and as a result the DIPSHIFT curvescoincide. In the fast limit, the molecular motion produces arandomization of the distribution of dipolar local fields, lead-ing to coinciding DIPSHIFT patterns when the order param-eter S is chosen appropriately. Note also that the initial partof all curves coincide, which arises from the fact that at smallevolution times, the AW approximation is valid even for non-

Gaussian frequency distributions17 �simply, the curve in thisregime is sensitive only to the second and not the highermoments of the distribution�. Thus, this result shows that theuse of analytical approaches based on the AW approximationwould be possible only for short evolution times for anyspinning speed, but it also points to a possibly better overallagreement at higher spinning frequencies �above 7 kHz forC–H groups�. Despite that, it is not possible to state that theDIPSHIFT curves obtained by spin dynamics simulationsand Eq. �1� will actually agree for intermediate motions.Hence, to investigate the limit validity of Eq. �1� in the in-termediate regime, a set of spin dynamics simulations forjumps among different numbers of sites was carried out andcompared with the results obtained via Eq. �1�.

Figure 4�b� shows a set of spin dynamics simulations ofthe DIPSHIFT curves considering 2-, 3-, 8-, and 16-sitejumps on a cone of aperture of 64° �see model in Fig. 2�. For3-, 8-, and 16-site jumps, all curves lie on top of each other,except for the 2-site jump process. For 3-, 8-, and 16-sitejumps, the only effect of increasing the number of sites with-out changing the motion geometry is a constant factor in therelation between the correlation time and the apparent mo-tional rates �C=1 / �Nsk�, which was taken into account inFig. 4�b�.20,37,38 Note also that the deviation between thecurves for two- and n-site jumps decreases for shorter t1

values and slower rates, but even in the short t1 part of thecurve, a deviation is still apparent. This is actually an ex-pected behavior because for two-site jumps, the average ten-sor in the fast limit is asymmetric, which is not the case ofmultisite jumps. Figure 4�c� shows a comparison betweenDIPSHIFT curves obtained by spin dynamics simulationsand Eq. �1�. Except for the rigid and fast limit cases alreadydiscussed, the agreement between the two approaches ispoor, in particular, at t1=TROT. However, at short t1 values�see right hand side of Fig. 4�c��, the agreement is actuallyfairly good. In fact, this can be understood by consideringthat at short evolution times, no Gaussian distribution of fre-quencies is needed for the AW approximation,16,17 whichmakes Eq. �2� or �3� good approximations even for finite-sitejump motions. This is an important point concerning the use-fulness of DIPSHIFT experiment in the intermediate regimebecause keeping the right limit a simple formula can be usedto quantify the data, providing a good estimate of the mo-tional correlation times. The case considered above might bea quite unfavorable case for testing agreement between thespin dynamics simulation and the AW based formula becausethe spin dynamics simulation considers a single C–H pair�Pake-like frequency distribution� executing jump process.Nevertheless, even for this very unfavorable case, acceptableagreement was found between both approaches when short t1

times are considered. Note that upon increasing the numberof spins �CH2 or CH3 groups�, the approach should be evenbetter because the stationary frequency distribution becomesmore “Gaussian-like.” This was actually confirmed by fur-ther simulations. Therefore, Eq. �1� seems to be a good ap-proximation for describing the short t1 values behavior of theDIPSHIFT curves even for jump motions and can be used fora quick estimation of the correlation times. A drawback ofusing Eq. �1� is that the apparent T2 decay is not well pre-

104505-6 deAzevedo et al. J. Chem. Phys. 128, 104505 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 7: Intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e., i it must permit the appli-cation of MAS and 1H decoupling during the acquisition

dicted by the formula, so this approach can only be used topredict the part of the DIPSHIFT curves where the C–Hcoupling averaging occurs. This reduces the dynamic win-dow of application to motions occurring with correlationtimes in the range of �20–0.1 �s. In contrast, if the spindynamics simulation is used, the dynamic window is ex-panded to between 1 ms and 0.1 �s. Note that this dynamicwindow is rather complementary to exchange experiments�1 s to 1 ms� and T1 relaxation measurements��10 ns�.

At this point, we would like to stress the conceptualsimilarity of the presently described experiments with whatbecame known as “dynamic MAS experiments.”18,21,22 Simi-lar to the well-known chemical exchange in liquids or thedynamic line shape in solids, the MAS spectra exhibit abroadening of the MAS side bands when the MAS rate be-comes comparable to the jump rate of the dynamic process.As emphasized by Suwelack et al.,16 the actual broadeningdepends on the geometry of the process and only a treatmentthat takes this into account can provide reliable kinetic pa-rameters from full fitting, which is to some extent the con-clusion obtained from the DIPSHIFT simulations discussedabove �a more complete discussion about such comparison isprovided in the supplementary material39�. However, in theDIPSHIFT experiment, there is a well-defined limit as towhen an analytical description can be used, which is an im-portant advantage. Another drawback of the experimentallymuch simpler line-shape experiments is the fact that the lineshape is not only governed by the interference of the molecu-lar dynamic process with the MAS rotation, but also by theheteronuclear dipolar decoupling26 that is much more diffi-cult to include in an analytical description or in the Floquettreatments. Qualitatively, in addition to the broadening thatappears when the jump rate is on the order of the MAS rate�typically some kilohertz�, another broadening sets in athigher temperatures when the molecular jump rate becomes

comparable to the frequency of the 1H decoupling, which isof the order of some tens of kilohertz. As it is shown below,this leads to the deviation of the calculated line shapes fromthe experimental data. Note, however, that in the spin dy-namics simulations of the DIPSHIFT, the LG decouplingwas taken into account explicitly and, as a result, this effectis part predicted of the analysis. Yet, the fact that spin dy-namics simulations and Eq. �1� agree at short t1 values alsoshows that the interference with the time scale of the LGdecoupling �the actual rf� is not too important. The reason isof course that DIPSHIFT is a constant-time approach, wherethe LG period is increased at the expense of normal cw de-coupling, which both operate at similar radio frequencies.

Analysis of the behavior of DIPSHIFT curvesfor complex motions

In the previous section, we have shown some generalcharacteristics of the DIPSHIFT curves. For that we usedsimplified motional models, usually involving single ratesand jumps between specific angles. However, in many casesmolecular motions in organic solids are much more complexand cannot be readily described by these simplified models.Two important features of these complex motions are thepresence of distributions of reorientation angles and correla-tion times, making it important to analyze the behavior of theDIPSHIFT curves in such cases.

The first topic to be addressed is the dependence of theDIPSHIFT patterns on the reorientation angle. For this weconsider two-site jump processes occurring with differentangles. Figure 5�a� shows the behavior of the DIPSHIFTcurves as a function of the reorientation angles for three mo-tional rates. In the slow-motion limit �Fig. 5�a� left�, the maineffect of increasing the reorientation angle is to decrease theintensity at t1=TROT, i.e., to reduce the effective T2 decay. Inthe intermediate regime �Fig. 5�a� middle�, there is a simul-taneous reduction of the effective coupling and effective T2

FIG. 5. �Color� �a� Spin dynamicssimulations of DIPSHIFT patterns atvarious reorientation angles with �left�k=104 Hz, �middle� k=5105 Hz,and �right� 5106 Hz. �b� Spin dy-namics simulations of DIPSHIFT pat-terns for distribution of reorientationangles: �left� k=104 Hz and �middle�k=5105 Hz. Ratio �right� betweenI�t1=TROT� / I�t1=TROT /2� vs .

104505-7 Intermediate motions in solid-state NMR experiments J. Chem. Phys. 128, 104505 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 8: Intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e., i it must permit the appli-cation of MAS and 1H decoupling during the acquisition

decay for increasing reorientation angles, and in the fast re-gime �Fig. 5�a� right�, only the effective coupling is reduced.

The usual variation that is under external control is achange in the motional rate �i.e., by varying the sample tem-perature�, and the increase in the motional rate can �or can-not� be accompanied by a change in the motional amplitude.The discrimination between these effects is generally diffi-cult, and the use of a combination of methods is usuallyrequired.40 However, the behavior of DIPSHIFT curves inthe intermediate regime is indeed sensitive to such a phe-nomenon. As just mentioned, in the intermediate-motion re-gime, the effect of increasing the reorientation angle on theDIPSHIFT curves is to simultaneously reduce the effectiveCH coupling and the intensity at t=TROT. Since this trend iscommon for any motional rate in this regime, it is also vis-ible when the rate and the reorientation angle increase simul-taneously. If we compare Fig. 5 �middle� and Fig. 3, whereno changes in the reorientation occur, we realize that theevolution of the DIPSHIFT patterns as a function of the rateshould be quite distinguishable. Thus, by monitoring thechanges in the shape of the DIPSHIFT patterns as a functionof temperature, it might be possible to identify this trend andassociate it with a simultaneous increase of the motionalrates and amplitude. However, this would always be just ahint, as the complexity of the motion or a change in themotion geometry might bias such a simple analysis. For acomplete elucidation of the process, more specific experi-ments that are able to probe separately the dynamic �corre-lation times� and the geometric part �reorientation angles�,e.g., exchange NMR experiments,40 should be used.

Another feature that can be addressed is the presence ofreorientation angle distributions. Figure 5�b� shows a set ofDIPSHIFT patterns for different motional rates considering aGaussian distribution of reorientation angles centered at 0°with various widths. The same general behavior as in thecase of just increasing single angle is observed but with asmaller scale of variation. Note that the wider the distribu-tion is, the smaller the intensity is at I�t1=TROT�, and thesmaller the apparent coupling is at I�t1=TROT /2�. This is bestobserved in a plot of the intensity ratio I�t1=TROT� / I�t1

=TROT /2� shown at the right side of Fig. 5�b�.Another usual feature found for complex motions is the

presence of a distribution of correlation times/rates. Figure 6shows simulated DIPSHIFT patterns for different widths ofcorrelation time distributions, where three-site jumps on acone with 54° opening angle are considered. Note that in theabsence of a correlation time distribution �Fig. 6�a��, the in-tensity at t1=TROT varies basically between 0 and 1, whilefor the case of a log-Gaussian distribution with a width oftwo decades �Fig. 6�b��, this variation range is considerablysmaller. It is actually rather straightforward to qualitativelyrealize the occurrence of a distribution. If the temperature-dependent data exhibit a smiley-like appearance, then a dis-tribution is very likely. This was also the case for a reorien-tation angle distribution, but note that in Fig. 6�b�, a decreasein the effective coupling with a simultaneous increase of theeffective T2 decay is observed. The quantitative treatmentrequires, of course, numerical calculations and comparisonswith the experimental data. As another demonstration, Fig.

6�b� shows the curve for k=105 Hz, i.e., for the case of maxi-mum dynamic broadening or minimum apparent T2, but withdifferent distribution widths. It is obvious that the narrowerthe distribution is, the shorter the T2 decay is at I�t1=TROT�and the smaller the apparent coupling is at I�t1=TROT /2�.This is again shown clearer by the ratio I�t1=TROT� / I�t1

=TROT /2�, which is on the order of 1 for no or a very narrowdistribution but increases with increasing distribution width.Note again that this behavior is contrary to the one observedfor a distribution of reorientation angles. Again, this can onlyserve as qualitative indicator, and a detailed analysis by nu-merical simulations is inevitable for quantitative results.

OTHER REMARKS

Despite the fact that the above approach seems to befeasible, there are two drawbacks that must be addressed.First, if cross polarization is used as the excitation method,the presence of a distribution of correlation times leads to abias in the fitted distribution because of the signal decayduring contact time. This holds, in particular, for segmentsdeep in the intermediate regime �with rates �105 Hz� �seethe supplementary material39�. To minimize the signal decay,it is vital to keep the contact time in the CP as short aspossible. Another possible procedure for minimizing the CPsignal decay is to perform a CP and contact time optimiza-tion for each experimental temperature. The second pointconcerns the remote 1H nuclei. The dipolar couplings be-tween the nuclear spin of a given carbon and the remote 1Hnuclear spin is indeed significantly smaller than for directlybonded spins. However, these remote spins are very abun-dant in organic systems, which may lead to significant ef-fects. In this sense, the two spin approach, as used here,might not be sufficient for a good description. Two cases canbe discussed when considering remote spins. If they arerigid, their presence will only add a constant contribution tothe DIPSHIFT curves. In this case they can actually helpwith the cross polarization, avoiding loss of signal due to

FIG. 6. Spin dynamics simulations of DIPSHIFT curves for a three-sitejump geometry considering the presence of distribution of correlation times.�a� Left: DIPSHIFT curves at various motional rates; right: DIPSHIFTcurves at various average motional rates considering a log-Gaussian distri-bution of rates with � =2 decades. �c� Left: DIPSHIFT curves consideringlog-Gaussian distributions of � =0, 0.5, 1, 1.5, 2, 3, and 4 decades at fixedmotional rates; right: relation between the DIPSHIFT intensities at t1

=TROT and t1=TROT /2.

104505-8 deAzevedo et al. J. Chem. Phys. 128, 104505 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 9: Intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e., i it must permit the appli-cation of MAS and 1H decoupling during the acquisition

intermediate regime motions during the CP, as discussedabove. In contrast, if they present some mobility, they mighthave some contribution to the shape of the DIPSHIFTcurves. However, it is expected that the effect of the remotespin will only appear for long evolution times, which alsopoints to the use of higher spinning frequencies in the DIP-SHIFT experiments in order to reduce the maximum evolu-tion times. Note that in experiments that uses recoupling,rotational echo double resonance �REDOR�-like experi-ments, the effect of remote spins is expected to be muchhigher because the evolution times are longer and the inter-action is “amplified.” In the case of the DIPSHIFT experi-ments, at least according to our simulations, the effect ofremote spins is not so important and two spin simulationsshould be adequate. In fact, the multispin scenario favors theapplication of the AW approximation, so the presence of re-mote spins also does.

EXPERIMENTAL RESULTS

Figure 7 compares experimental data versus temperaturefrom imidazole methyl sulfonate together with spin dynam-ics simulations. The agreement is quite remarkable. By plot-ting the jump rates in an Arrhenius plot, we can extract anactivation energy of 70�5 kJ /mol, which again is in goodagreement with previously obtained data.31,34 However, themelting of this compound prevents obtaining of the secondpart of the dephasing curves as shown in Fig. 7, i.e., the

decrease of the apparent dipolar coupling. To cover the fulldynamic range of the method, we turned to TMSI. Here, thethree methyl groups are arranged in a tetrahedral symmetryand perform threefold jumps. The fast methyl rotation leadsto a preaveraging of both the CSA as well as the dipolartensor. Figure 8 shows temperature-dependent MAS spectratogether with corresponding simulations. Note the discrepan-cies at higher temperatures, which are due to the interferenceof the molecular motion in the kilohertz range with the co-herent 1H irradiation, which is another source of dynamicline broadening from the deterioration of the 1H-decouplingefficiency. This effect is not included in the line-shape simu-lation, and this is, as mentioned above, one of the crucialdrawbacks of the dynamic MAS approach.

Figure 9 �top� shows experimental and simulatedtemperature-dependent dephasing curves. In Fig. 9 �bottom�,the correlations times �C= �3k�−1 are plotted in an Arrheniusdiagram, from which an activation energy of about76 kJ /mol is extracted. As comparison, previous 2H line-shape data �see Fig. 5 from Ref. 12� are plotted, too. It can beseen that the two experimental methods yield similar dataand that the DIPSHIFT is also sensitive to longer correlationtimes. This is expected as it is more sensitive for motionswith rates of the order of the dipolar coupling �a few kilo-

FIG. 7. Experimental and simulated DIPSHIFT curves vs temperature andthe corresponding Arrhenius plot for imidazole methyl sulfonate.

FIG. 8. Experimental and simulated MAS spectra of TMSI as a function oftemperature.

104505-9 Intermediate motions in solid-state NMR experiments J. Chem. Phys. 128, 104505 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 10: Intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e., i it must permit the appli-cation of MAS and 1H decoupling during the acquisition

hertz in this case�, while for the 2H line shape, the width ofthe 2H spectrum �about 40 kHz� is the determining factor.Note that while the DIPSHIFT works in natural abundance,the 2H experiments require isotopic enrichment which is notfeasible in all cases. Another comparison that is displayed inthe Arrhenius plot is the correlation times obtained by thefitting of the short t1 range of the experimental data�t1 /TROT�0.3� by Eq. �1� �AW approach�. The agreement isindeed fairly good, showing that this approach can also beused. However, it should be mentioned that there is consid-erable error in the correlation times extracted from this ap-proach in Fig. 9, which is mostly due to the poor sampling ofour data for t1 /TROT�0.3. Actually, one important experi-mental detail of the DIPSHIFT experiment is that the sam-pling of the t1 points should coincide with the LG period �2�proton pulse during LG irradiation�, which reduces the num-ber of t1 points.

Finally, we turn to the photoluminescent polymer MEH-PPV, which was already the subject of NMRinvestigations.32,33 It was shown that the � relaxation, whichis related to the optical and electric properties of the materialis, due to the side-chain motion. Around room temperature,this motion is in the fast limit and, therefore, the CH carbonlabeled as 11 in Fig. 2 is an ideal candidate for acquiringdetailed dynamic information on this motion. The NMRspectrum of MEH-PPV as a function of temperature in Fig.10�a� shows 13C CP/MAS temperature-dependent spectra ofMEH-PPV. The line assignment �according to the chemicalstructure shown in Fig. 2� is given in the spectrum acquiredat 353 K. It is obvious for carbon 11 that there is atemperature-dependent broadening of the resonances belong-

ing to the side-chain carbons, while those of the main chainremain fairly constant, confirming a side-chain specific pro-cess that reaches the intermediate regime at lower tempera-tures. The broadenings can in principle be simulated usingthe procedures mentioned above. However, the very densespectrum and the small degree of broadening would makethis very ambiguous. Also, the knowledge about the principalCSA values as well as the orientations of the CSA tensors inthe molecules would be necessary. A much more elegant ap-proach is the DIPSHIFT that requires fitting of the dephasingcurves for the different resonances, for which only the ex-perimental line intensities are required. The use of the C–Hdipolar coupling, for which the principal values and orienta-tions are well known, makes it also much easier to interpret,as compared to the CSA interaction that needs to be simu-lated in the cases of dynamic MAS experiments. Also, fromour experience with dynamic MAS experiments, we are con-fident that it would be impossible to perform an unambigu-ous simulation like the one presented below. In particular, thedistribution of �C would be difficult to determine withoutprior knowledge. Of course, also for DIPSHIFT simulationsan a priori knowledge of the motion geometry is desired. Inthe case of MEH-PPV, previous measurements32 have shown�using DIPSHIFT experiments� that the CH2 group attachedto the main chain �labeled 10 in Fig. 3� is rather rigid �notemperature dependence was detected in the DIPSHIFTcurves of this group from −50 to 50 °C�, while the CHgroup 11 is already mobile in this temperature range. This is

FIG. 9. Experimental and simulated DIPSHIFT curves vs temperature andthe corresponding Arrhenius plot for TMSI. FIG. 10. �a� Experimental MAS spectra of MEH-PPV as a function of

temperature. �b� Right: experimental DIPSHIFT curves for MEH-PPV at afixed average rate and various widths of Gaussian distributions of rates; left:experimental DIPSHIFT curves for MEH-PPV vs temperature and the cor-responding simulations considering two decades of rate distribution.

104505-10 deAzevedo et al. J. Chem. Phys. 128, 104505 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 11: Intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e., i it must permit the appli-cation of MAS and 1H decoupling during the acquisition

largely consistent with a motional model were the CH group11 executes n-site jumps around the H2C10– 11CH bond,which suggests a model of n-site jumps on a cone with anopening defined by the relative orientation of the CH bondwith respect to the C–C bond �ideally 109°�. Note that thenumber of jump sites is undetermined, but as shown in Fig.4, for symmetric jumps over more than two sites, all theDIPSHIFT curves lie on top of each other. Thus, for the sakeof simulation time, we just consider three-site jumps. Thepossibility of a two-site jump was actually excluded by thevalue of the order parameter extracted from the DIPSHIFTcurve at high temperature �fast limit�.

Figure 10�b� exhibits the experimental and simulateddata obtained for the CH carbon 11. Figure 10�b� �left� showsthe experimental MEH-PPV DIPSHIFT curve obtained at−20 °C and a set of simulated curves for an average rate of�K�=105 Hz and Gaussian rates distributions with 0.0, 0.5,1.0, 2.5, 2.0, 3.0, and 4.0 decades. As it can be observed, it isnot possible to simulate the experimental data using a singlecorrelation rate. It rather prompts for the assumption of acorrelation time distribution of 2.0 decades, as it is fre-quently found in polymers.37,41 As discussed before, the be-havior of the DIPSHIFT curves as a function of temperaturealso point to that. Figure 10�b� �right� shows experimentalDIPSHIFT curves as a function of temperature. These curvescould be well fitted by threefold jump with fixed geometrybut with a log-Gaussian rate distribution of about two de-cades, centered at different �K� values. No corrections toaccount for the loss of CP signal due to intermediate motionswere applied. This is justified by the presence of many rigidprotons in the vicinity of the CH group 11, which reduces thebias CP effect substantially. Further experiments on this ma-terial, in particular, with the extension to slower motions andthe application of exchange NMR methods, are in progressand will be presented elsewhere.

As a final point, we would like to place a comment onpossible extensions and alternatives. The approach as de-scribed above might be improved in two directions. First,there is always the value of the intrinsic T2 decay that arisesfrom imperfections of the LG-building block. This needs tobe determined separately on a rigid sample. Second, whendealing with methyl groups, there is already a preaveragingof the dipolar coupling which reduces the variation range ofthe normalized intensities. The former would ask for aconstant-time version of the experiment, while the latter canbe solved by a version that enhances the effect of existingdipolar coupling in the experiment. Both requirements arefulfilled by the DIPSHIFT version introduced by Hong etal.11 The theory presented here is applicable for this methodtoo; however, the simulation program of course has to bematched to the actual pulse sequence. Note that there is alsoan AW theory for recoupling of dipolar interactions that aredescribed in Ref. 20 and can, in principle, be used, but thelimits should be checked as done here.

CONCLUSIONS

We extended the application of the DIPSHIFT-NMRmethod, an SLF method that was designed to be applied in

the fast-motion limit only, toward the quantification ofintermediate-range motions. For that, a more elaborated nu-merical calculation approach is necessary; however, for spe-cial cases, the correlation time of motion can also be ex-tracted from a simple formula based on the Anderson–Weissapproach and, thus, a truly model-free approach to obtain thekinetic parameters is provided. In conjunction with that, wehave also discussed the signal loss due to the presence ofintermediate regime motions during the cross-polarizationtransfer, and some strategies to avoid this loss are proposed.Although the NMR experiment itself is well known, its com-bination with the presented numerical routine opens up anew range of applications, overcoming some of the problemsof line-shape experiments that had to be used so far. In fact,this approach closes the gap between solid-state exchangeNMR experiments, which are sensitive to slow motions, andthe original DIPSHIFT �or more general SLF� experimentsfor the fast processes.

ACKNOWLEDGMENTS

The authors acknowledge the Deutsche Forschungsge-meinschaft �DFG, SFB 418� and DAAD Brazil-Germanyprogram for financial support. E.R.dA., A.A.S., and T.J.B.thank CNPQ, CAPES, and FAPESP for the financial support.

1 E. R. Andrew and R. G. Eades, Proc. R. Soc. London, Ser. A 216, 398�1953�.

2 E. Andrew, J. Chem. Phys. 18, 607 �1950�.3 M. Hong, X. Yao, K. Jakes, and D. Huster, J. Phys. Chem. B 106, 7355�2002�.

4 R. Hester, J. Ackerman, B. Neff, and J. Waugh, Phys. Rev. Lett. 36, 1081�1976�.

5 M. Munowitz, R. Griffin, G. Bodenhausen, and T. Huang, J. Am. Chem.Soc. 103, 2529 �1981�.

6 D. Cory and W. Ritchey, Macromolecules 22, 1611 �1989�.7 B. Van Rossum, C. De Groot, V. Ladizhansky, S. Vega, and H. De Groot,J. Am. Chem. Soc. 122, 3465 �2000�.

8 P. Tekely, D. Nicole, J. Brondeau, and J. Delpuech, J. Phys. Chem. 90,5608 �1986�.

9 N. Zumbulyadis, Phys. Rev. B 33, 6495 �1986�; K. Schmidt-Rohr, J.Clauss, and H. Spiess, Macromolecules 25, 3273 �1992�.

10 H. J. M. D. G. A. C. Kolbert, M. H. Levitt, M. G. Munowitz, J. E.Roberts, G. S. Harbison, J. Herzfeld, and R. G. Griffin, MultinuclearMagnetic Resonance in Liquids and Solids �Kluwer, Dordrecht, 1990�.

11 M. Hong, J. Gross, and R. Griffin, J. Phys. Chem. B 101, 5869 �1997�.12 D. Reichert, Annu. Rep. NMR Spectrosc. 55, 159 �2005�.13 A. Bain, Biochem. Cell Biol. 76, 171 �1998�.14 H. Spiess and H. Sillescu, J. Magn. Reson. �1969-1992� 42, 381 �1981�;

R. Vold, R. Vold, and N. Heaton, Adv. Magn. Reson. 131, 381 �1989�.15 A. Schmidt and S. Vega, J. Chem. Phys. 87, 6895 �1987�.16 D. Suwelack, W. Rothwell, and J. Waugh, J. Chem. Phys. 73, 2559

�1980�.17 P. Anderson and P. Weiss, Rev. Mod. Phys. 25, 269 �1953�.18 S. Clough and K. W. Gray, Proc. Phys. Soc. London 79, 457 �1962�.19 D. Fenzke, B. Gerstein, and H. Pfeifer, J. Magn. Reson. 98, 469 �1992�.20 J. Hirschinger, Concepts Magn. Reson., Part A 28a, 307 �2006�.21 A. Schmidt, S. Smith, D. Raleigh, J. Roberts, R. Griffin, and S. Vega, J.

Chem. Phys. 85, 4248 �1986�.22 D. Reichert, G. Hempel, H. Zimmermann, H. Schneider, and Z. Luz,

Solid State Nucl. Magn. Reson. 18, 17 �2000�; Z. Luz, R. Poupko, and S.Alexander, J. Chem. Phys. 99, 7544 �1993�.

23 M. Duer, J. Magn. Reson., Ser. A 119, 204 �1996�.24 J. Long, B. Sun, A. Bowen, and R. Griffin, J. Am. Chem. Soc. 116,

11950 �1994�; J. Kristensen, H. Bildsoe, H. Jakobsen, and N. Nielsen, J.Magn. Reson. �1969-1992� 92, 443 �1991�; M. Hologne and J. Hirsch-inger, Solid State Nucl. Magn. Reson. 26, 1 �2004�.

25 K. Saalwachter and I. Fischbach, J. Magn. Reson. 157, 17 �2002�.

104505-11 Intermediate motions in solid-state NMR experiments J. Chem. Phys. 128, 104505 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 12: Intermediate motions as studied by solid-state separated ... · abundance in organic solids, i.e., i it must permit the appli-cation of MAS and 1H decoupling during the acquisition

26 W. Rothwell and J. Waugh, J. Chem. Phys. 74, 2721 �1981�.27 K. Takegoshi and K. Hikichi, J. Chem. Phys. 94, 3200 �1991�.28 M. Lee and W. Goldburg, Phys. Rev. 140, 1261 �1965�.29 A. Bielecki, A. Kolbert, and M. Levitt, Chem. Phys. Lett. 155, 341

�1989�.30 E. Vinogradov, P. Madhu, and S. Vega, Chem. Phys. Lett. 314, 443

�1999�.31 I. Fischbach, H. W. Spiess, K. Saalwachter, and G. R. Goward, J. Phys.

Chem. B 108, 18500 �2004�.32 R. F. Cossiello, E. Kowalski, P. C. Rodrigues, L. Akcelrud, A. C. Bloise,

E. R. deAzevedo, T. J. Bonagamba, and T. D. Z. Atvars, Macromolecules38, 925 �2005�.

33 A. C. Bloise, E. R. deAzevedo, R. F. Cossiello, R. F. Bianchi, D. T.Balogh, R. M. Faria, T. D. Z. Atvars, and T. J. Bonagamba, Phys. Rev. B71, 174202 �2005�.

34 G. Goward, K. Saalwachter, I. Fischbach, and H. Spiess, Solid StateNucl. Magn. Reson. 24, 150 �2003�.

35 J. Vanvleck, Phys. Rev. 74, 1168 �1948�.

36 J. Roland and D. Michel, Magn. Reson. Chem. 38, 587 �2000�; P. Ma-gusin, V. Zorin, A. Aerts, C. Houssin, A. Yakovlev, C. Kirschhock, J.Martens, and R. Van Santen, J. Phys. Chem. B 109, 22767 �2005�.

37 K. Schmidt-Rohr and H. W. Spiess, Multidimensional Solid-State NMRand Polymers, 1st ed. �Academic, London, 1994�.

38 A. Abragam, Principles of Nuclear Magnetism, 1st ed. �Oxford Univer-sity Press, New York, 1961�.

39 See EPAPS Document No. E-JCPSA6-128-026805 for more details ondynamic MAS experiments and effect of intermediate regime motions inthe cross-polarization CP under MAS. This document can be reachedthrough a direct link in the online article’s HTML reference section or viathe EPAPS homepage �http://www.aip.org/pubservs/epaps.html�.

40 H. Kaji, N. Miura, and K. Schmidt-Rohr, Macromolecules 36, 6100�2003�.

41 M. Cook, D. Watts, and G. Williams, Trans. Faraday Soc. 66, 2503�1970�; G. Williams, J. Non-Cryst. Solids 131, 1 �1991�; F. Cumbrera, F.Sanchezbajo, F. Guiberteau, J. Solier, and A. Munoz, J. Mater. Sci. 28,5387 �1993�; J. Phillips, Rep. Prog. Phys. 59, 1133 �1996�.

104505-12 deAzevedo et al. J. Chem. Phys. 128, 104505 �2008�

Author complimentary copy. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp