Solid State NMR Basic Principles and Practice
Solid State NMR Basic Principles and Practice
SOLID-STATE NMR:
Logo.eps
Solid-State NMR: Basic Principles & Practice
Copyright © Momentum Press®, LLC, 2012.
DOI: 10.5643/9781606503522
10 9 8 7 6 5 4 3 2 1
PREFACE
ABOUT THE AUTHORS
1 INTRODUCTION
1.1 The utility of NMR
1.2 A preview of solid-state NMR spectra
1.3 The solid state
1.3.1 Introduction
1.3.2 Symmetry in the crystalline state
1.3.3 Effects of crystal structure on NMR
1.3.4 Types of solids
1.4 Polymorphism, solvates, co-crystals & host:guest systems
1.5 NMR of solids & the periodic table
6 QUADRUPOLAR NUCLEI
6.1 Introduction
6.2 Characteristics of first-order quadrupolar spectra
6.3 First-order energy levels & spectra
6.4 Second-order zero-asymmetry cases
6.4.1 Transition frequencies
6.4.2 Central-transition spectra: Static samples
6.4.3 Central-transition spectra: Rapid sample spinning
6.5 Spectra for cases with non-zero asymmetry: Central
transition
6.6 Recording one-dimensional spectra of quadrupolar nuclei
6.6.1 Nutation
6.7 Manipulating the quadrupolar effect
6.7.1 Variable-angle spinning
6.7.2 Double rotation
6.7.3 Dynamic-angle spinning
6.7.4 Multiple quantum magic-angle spinning
6.7.5 Satellite transition magic-angle spinning
6.7.6 Summary for spectroscopy of half-integer quadrupolar
nuclei
6.8 Spectra for integral spins
7 RELAXATION , EXCHANGE & QUANTITATION
7.1 Introduction
7.2 Relaxation
7.2.1 The cause of relaxation
7.2.2 Proton relaxation times
7.2.3 Lineshapes & linewidths for non-spinning samples
7.2.4 Relaxation & high-resolution measurements combined
7.2.5 Measuring relaxation times
7.3 Exchange
7.3.1 Positional exchange
7.3.2 Hydrogen exchange
7.3.3 Reorientation without a change in isotropic chemical
shift
7.3.4 Diffusive motion
7.3.5 “Soft” solids
7.3.6 Interference
7.4 Quantitative NMR
7.4.1 Relative intensity
7.4.2 Absolute intensity
7.5 Paramagnetic systems
7.5.1 Relaxation effects
7.5.2 Shift effects
APPENDICES
INDEX
PREFACE
Nuclear magnetic resonance (NMR) has proved to be a uniquely powerful
and versatile spectroscopy, and no modern university chemistry department
or industrial chemistry laboratory is complete without a suite of NMR
spectrometers. The phenomenon of nuclear spin may seem an odd basis for
an analytical tool, but it is the relative isolation of the nuclear spin from its
surroundings that makes it an ideal noninterfering probe of the electronic
environment. Different sites are clearly identified by their chemical shifts,
while J couplings in 1 H spectra provide connectivity information. The
combination of these two complementary interactions, plus the formidable
array of different NMR experiments developed since the arrival of Fourier
transform NMR in 1966, has revolutionized the practice of chemistry.
While the original discovery of NMR involved both solids and liquids, the
application of NMR to materials in solid form developed at a markedly
slower rate than its solution-state counterpart until relatively recently. The
most obvious explanation for this difference is the fact that molecular
mobility in isotropic solutions averages anisotropic interactions (such as
shielding) to their isotropic values. In particular, for solids it is necessary to
consider a number of NMR interactions (notably the dipolar and
quadrupolar interactions) that can generally be ignored for the solution
state. These can significantly reduce spectral resolution and complicate
interpretation of spectra. On the other hand, the direct effect of these
interactions means that solid-state NMR is potentially a much richer
information source than NMR of solutions.
A second, perhaps less appreciated, reason for the relatively tardy progress
of solid-state NMR is the nature of the solid state itself. Solids, and
especially the samples the solid-state NMR spectroscopist is asked to deal
with, are rarely the simple monocrystals beloved by diffraction
crystallographers. Solid-state NMR may well give broad, featureless lines
for amorphous or heterogeneous samples, but this reflects the underlying
nature of the system—the chemical shift of a given nucleus may be a
distribution and not a single value as it would be if the sample were
dissolved. As a result, solid-state NMR may be providing too much
information, so that the art of solid-state NMR spectroscopy lies in finding
the right approach to refining the information content.
It is this complex interaction between the nature of the sample and the
multiple anisotropic NMR interactions that makes solid-state NMR a
challenging technique to master, and the treatments in undergraduate texts
do little to dispel the image of a dark and troublesome technique. The
purpose of the current text is to provide a bridge between the familiar world
of isolated molecules relevant to solution-state NMR and the subtle world
of solid materials. We hope that this introduction and survey will be of
value to both established researchers wishing to learn what solid-state
NMR can (and cannot) do for their systems and graduate students starting
work in this area. We have deliberately avoided detailed mathematical
treatments in the early chapters, concentrating on providing that essential
qualitative “feel” for the different aspects of solid-state NMR. The
theoretical treatment, which will be of most relevance to those starting a
career in the field, is delayed to chapter 4. This survey of the theoretical
tools used in solid-state NMR should allow the reader to tackle the more
specialized literature with increased confidence. The advanced experiments
described in chapters 5–8 can only be completely understood in terms of
the underlying theory, but again a basic feel for the experiments and what
information they provide should still be clear without such knowledge.
INTRODUCTION
NMR spectra are generally obtained from samples in strong magnetic fields
by recording the response to radiofrequency (RF) radiation, which is
normally applied in the form of short pulses. The detailed behavior of the
nuclear spins depends on the RF pulsing regime, on the spin properties, and
on the interactions of the spins. Thus, one must evaluate the terms
governing the energy of the spins in the magnetic field and subject to
irradiation. The Zeeman effect of the spin magnets in the applied magnetic
induction field (B 0 —conventionally in the z direction) almost always
forms by far the largest contribution to the energy. The full Zeeman energy,
E Z , involves a sum over all the nuclear species in the sample. For a given
spin j (ignoring shielding), it is expressed as:
CH_01-01.wmf 1.1
CH_01-02.wmf 1.2
Figure 1.1. Carbon-13 spectra of alanine in the solid state (upper trace) and
in solution (lower trace). The latter was obtained using proton decoupling,
as is usual for solutions. The former involved no special techniques.
ch01-F02.ai
1.3.1 INTRODUCTION
The principal way in which solids differ from fluids lies in the relative lack
of molecule-level motion. In isotropic solution, molecules and ions tumble
rapidly, randomly, and chaotically. Therefore, any properties that depend on
molecular orientation become averaged to their isotropic values. The
internal NMR interactions of importance come into this category. Such
interactions each involve products of two vectors, so they are classed as
tensors (see chapter 2 ). In fact, two of the most important features, namely
the electric quadrupolar and magnetic dipolar interactions, are averaged to
zero by isotropic tumbling and therefore do not affect the NMR transition
frequencies of solutions. Simple interpretations of spectra can ignore their
existence, though they do affect relaxation times. Moreover, the other two
important interactions (shielding, which gives rise to chemical shifts, and
indirect coupling) are simplified so that for solutions a single scalar
chemical shift results for each chemical site and a scalar coupling constant
(symbol J ) exists for every pair of nuclear spins.
There are many types of solids and solid systems. The simplest (and most
commonly met with in chemistry research) is the crystalline state. In
“ideal” crystalline samples, the atoms (whether in molecular or in
framework solids) are effectively locked into positions. The structures as a
whole are described by unit cells wherein the atomic positions can be
determined by diffraction methods. The orientation of a crystal in B 0 will
determine the resonance frequency for each NMR-active nucleus.
ch01-F03.ai
ch01-F04.ai
Figure 1.4. Left: The two independent molecules (labeled U and V) in the
asymmetric unit of the a form of the steroid testosterone. These have
different environments and slightly different geometries. Right: Carbon-13
spectrum of α -testosterone (high-frequency region only), obtained under
conditions of proton decoupling and magic-angle spinning (see section 1.5)
to show crystallographic splittings. The asterisk indicates a spinning
sideband (see section 2.7).
For a static single crystal, atomic sites which are in identical chemical
positions will, in general, be in different orientations with respect to B
0 and will therefore give rise to different resonance frequencies. Then
more peaks are obtained than for the solution state, forming
crystallographic splittings . However, such splittings may be
eliminated by the technique of magic-angle spinning (MAS),
discussed later (section 1.5).
Independent molecules in an asymmetric unit will differ in both their
intermolecular (packing) and intramolecular (geometry) arrangements.
Therefore the environment of chemically analogous spins in (say) two
independent molecules will be different, and so they will give rise to
different resonance frequencies—another (more commonly observed)
cause of crystallographic splittings. This type of splitting cannot be
eliminated by MAS. A partial 13 C spectrum of α-testosterone is
shown in figure 1.4 to illustrate this phenomenon. It is difficult to
disentangle experimentally the influences of packing and geometry on
NMR spectra.
Moreover, a molecule that has symmetry in its isolated state (as in
solution) may well lose its symmetry in the crystalline state. Thus, for
instance, if the two sides of a paraphenylene group (related by
symmetry in the isolated molecule) have different local environments
in the crystal, the relevant spins will have different chemical shifts. In
such a case, the whole paraphenylene group is (part of ) the
asymmetric unit, though if the symmetry of the isolated molecule is
retained, the asymmetric unit may be half the molecule (plus any atom
on the effective mirror plane or symmetry axis relating the two sides
of the ring). The two situations are illustrated schematically in figure
1.5 .
The pharmaceutical drug barbital provides an example (figure 1.6 ). In
this case the intramolecular lack of symmetry induced by the
molecular environment in the crystal also extends to the two ethyl
groups in some forms. The structure of polymorphic form III (figure
1.6 (a)) consists of molecular chains in which the two halves of the
ring are related by symmetry, as are the two ethyl groups, so the
asymmetric unit is half a molecule. This means that only single
resonances are seen for the methyl and methylene carbons (as would
be the case for a solution-state spectrum). However, form I has a
double-chain structure (figure 1.6 (b)) in which there is no such
relationship, so an entire molecule is the asymmetric unit. Thus the
methyl and methylene signals are split into two. Note, however, that
the ring carbon signal is unsplit.
ch01-F05.ai
ch01-F06.ai
(a) The acquisition time to record a single free-induction decay (from tens
of milliseconds to several seconds; see inset 3.2). It is this range that is
generally referred to as the NMR timescale . If molecules are essentially
static (apart from vibrational motion) over such times, then the spectrum
will be a superposition of those for all molecular positions (e.g.,
conformations) present. However, if there is rapid motion, that is over times
shorter than this, an averaged spectrum will be obtained (see section 7.3).
Typically, the critical motional rates are those comparable to chemical shift
differences between corresponding nuclei for the different molecular
positions.
(b) The total accumulation time taken to record a spectrum, which may be
from minutes to hours or even days, depending on the number of free
induction decays that are recorded. If the compound is stable over this time,
its spectrum will be uniquely obtained. If reaction occurs, the spectrum will
be a superposition of all the compounds present.
There are a number of types of solid state that are not crystalline or not
perfectly crystalline. Some solids exhibit disorder. This may take a number
of forms; in particular, it may be static or dynamic (spatial or temporal
disorder). In either situation, diffraction techniques report occupancy
factors at the possible sites, implying that there are fractional atoms there.
In the former case, adjacent “unit cells” may have some atoms in different
positions, which do not change with time. In the latter case, there may be
exchange (over a particular timescale) of atoms between two or more
possible positions for any given unit cell; the rate of exchange will
determine the appearance of the NMR spectrum (see section 7.3). An
example is the 3:2 complex of phenol and triphenylphosphine oxide
(TPPO). A partial structure is shown in figure 1.7 . The phenol molecule is
rotationally disordered (and was thus solved in a CH_In_01-01.wmf
space group), with arrows indicating two notional oxygen half-atoms. With
this molecule fixed in one of its two arrangements, the phosphorus atoms of
the two TPPO molecules are nonequivalent. At the lowest temperature this
is the case: the molecules are static on the NMR timescale, so two 31 P
signals appear. However, as the temperature increases, rotation of the
phenol molecule becomes rapid, causing the signals to merge because an
average situation is observed by NMR. The NMR proves that the disorder
is dynamic (temporal, not spatial).
ch01-F07.ai
Figure 1.7. Left: Partial structure of the unit cell contents of a 3:2 adduct of
phenol and triphenylphosphine oxide. The other two phenol molecules in
the unit cell, which are not disordered, are omitted for clarity. Right:
Phosphorus-31 spectra as a function of temperature.
The different sites in disordered systems do not need to occur with rational
occupancy ratios. Many inorganic crystalline systems exist in which there
is a fractional excess of some types of atom (over and above the nominal
stoichiometric amount). In other cases, there is a deficiency of some types
of atom. Such systems are known as defect structures, and their nature has
consequences on the NMR spectra (e.g., on relative intensities of
resonances).
ch01-F08.ai
A final type of physical system that needs special consideration is the liquid
crystalline state, which can be considered as a half-way house between the
isotropic solution and crystalline solid states. Such systems have
considerable mobility analogous to that of a liquid but possess order
analogous to that of a solid in some dimensions. The behavior of liquid
crystals in NMR experiments is a specialized topic which will not be
discussed in this book. Suffice it to say here that dipolar and quadrupolar
interactions are not averaged to zero for liquid crystals and therefore
influence the resonance frequencies.
ch01-F09.ai
ch01-F10.ai
In many cases, the basic structure of the host compound contains cages or
channels that can be occupied by a range of guest molecules. Zeolites, for
instance, often contain both cages and channels, which result from the use
of organic molecules as templates during synthesis. Another example is
provided by the complexes of small organic molecules with urea. The urea
provides a host structure (unstable in the absence of guest molecules!)
containing linear tunnels into which elongated molecules such as alkanes
and their simple derivatives can fit. However, the repeat distance of the
urea host in the tunnel direction is not in general the same as the repeat
distance of the guest molecules, which is governed by their length. Such
structures fall into the class of incommensurate systems mentioned in
section 1.3.4. Frequently, the guest molecules in such complexes are highly
mobile in the host tunnels and high-quality spectra can therefore be readily
ob tained (figure 1.11 ). Diffraction te chniques struggle to precisely locate
the guest molecules in the tunnels because of their high degree of dynamic
disorder.
ch01-F11.ai
Figure 1.11. Top left: Structure of the inclusion compound between urea
and 1-fluorononane as determined by X-ray studies at ambient temperature,
viewed down the tunnels formed by the urea host. Note that the disordered
guest molecules do not give rise to Bragg scattering and so the tunnels
appear to be empty in the reported structure. Bottom: The low-frequency
part of the 13 C spectrum (showing the signals arising from the guest
molecule only), obtained under conditions of high-power proton
decoupling and magic-angle spinning (see section 1.5).
Chiral molecules may form crystals that each contain only one enantiomer,
but in such cases usually as a mixture of crystals of the two forms (a
conglomerate ). The NMR spectra of the two pure enantiomers (and of the
corresponding conglomerate) will be identical. Alternatively, crystals of an
internal racemate may form, with the enantiomers in a unit cell related by a
center or a plane of symmetry. The spectrum of a racemate will in general
differ from that of the related enantiomers, thus providing a powerful way
of distinguishing between the two situations.
As discussed above, the nature of the crystalline state implies that NMR
spectra will be more complex than those of isotropic solutions. Additional
considerations arise when microcrystalline powders are examined by NMR
(as is the normal case), since each crystallite will be separately oriented in
the applied magnetic field. This results in powder-pattern spectra for static
samples, as illustrated earlier. Three particular techniques, which are
discussed in more detail later, are of considerable importance in obtaining
high-quality, well-resolved spectra, especially from powdered samples.
Two of these relate to improvements in resolution, while the third is
concerned with increasing signal intensities:
Cross polarization (CP) . The third problem, which affects signal intensities
rather than resolution, is that relaxation times for dilute spins in solids tend
to be long (generally tens to hundreds of seconds, but in some cases as
much as hours), rendering the usual pulse-and-acquire technique for
recording free induction decays very inefficient (i.e., requiring long inter-
pulse delays; see section 3.3.1). This difficulty is ameliorated by the use of
CP from protons to the dilute spins (achieved by a pulse sequence as
described in section 3.4).
Quadrupolar nuclei with integral spin quantum numbers . Such nuclei are in
principle the most difficult to deal with, since they have no central
transitions. Fortunately, there are very few such nuclides, the principal ones
being 2 H, 6 Li, 10 B, and 14 N. In each case, there is an alternative
nuclide. However, it is sometimes valuable to have spectra of one or other
of these four (though not normally of 10 B). In fact, the quadrupole
moments of 2 H and 6 Li are small, so their spectra present no problem and
only 14 N can be said to still represent a difficulty (see section 2.6).
FURTHER READING
“Crystalline solids ”, D. McKie & C. McKie, Thomas Nelson & Sons Ltd.
(1974), ISBN 0 17 761001 8.
“Fundamentals of crystallography ”, Ed. C. Giacovazzo, Oxford University
Press (1992), ISBN 0 19 8555 79 2.
“Solid state chemistry ”, L. Smart & E. Moore, Chapman & Hall (1992),
ISBN 0 412 40040 5.
“Basic solid-state chemistry ”, A.R. West, John Wiley & Sons Ltd. (1988),
ISBN 0 471 91797 4.
“NMR crystallography ”, Eds. R.K. Harris, R.E. Wasylishen & M.J. Duer,
John Wiley & Sons Ltd. (2009), ISBN 978 0 470 69961 4.
NOTES
The transition between these states is the basis of NMR spectroscopy. The
energy difference between them is
2.2 TENSORS
Since vectors have three components each, any tensor R (in its Cartesian
representation) involves a 3 × 3 matrix, that is, nine components:
In other words, the shielding field is not in general parallel to the applied
field (figure 2.1 ).
However, tensors are often symmetric, that is, R ji = R ij and, even when
this is not the case, the antisymmetric contributions have little influence on
NMR spectra and are usually ignored. In such circumstances, there are only
six distinct components to a tensor, and it is possible to choose axes
(forming the principal axis system, PAS) in which R is diagonal:
The relation between them is missing image file Either definition may be
used for shielding. However, ΔJ is the normal form for indirect coupling.
This choice looks a little odd until it is realized that the X , Y , and Z
components are in alphabetical order in the spectrum (see figure 2.2 ),
either R ZZ ≥ R YY ≥ R XX or R ZZ ≤ R YY ≤ R XX (depending on the
system studied).
The following sections give details of the four tensor properties most
important for solid-state NMR, namely shielding, indirect coupling, dipolar
coupling, and quadrupolar coupling. Table 2.1 gives some of the general
properties of these tensors. The equations in this book are written such that
the coupling parameters (and their anisotropies) appear as frequencies
(other texts use angular frequencies).
Figure Missing
2.3 SHIELDING
In the more general case, there will be three different principal components
of the shielding tensor, as mentioned in section 2.2 above, and the powder
pattern will have, correspondingly, three turning points as in the three lower
spectra in figure 2.2 . The relevant equation for the effective shielding (to
be used in equation 2.18 for deriving energy-level diagrams and hence
transition frequencies) will be:
where θ j is the angle between the shielding principal axis j and the
magnetic field. This equation may be rewritten in terms of the anisotropy, ζ
, and asymmetry, η , as:
For powder patterns of the types shown in figure 2.2 , the turning points
give the values of the principal components (but not the orientation of the
PAS in the molecular/crystallographic frame). To obtain reasonably
accurate data, a bandshape-fitting procedure is required, as discussed in
section 8.2.1. However, in situations of nearly axial shielding,
differentiation of σ XX from σ YY is difficult, that is, values of η below 0.2
are not determined accurately.
Shielding anisotropies are rather small for 1 H, but are up to ∼ 200 ppm for
13 C and can be massive for some metallic nuclides.2 Their values,
together with those for shielding asymmetries, can be used to assign
resonances to chemical sites. They also give important crystallographic
information and frequently distinguish signals far better than do isotropic
shifts. They help to understand electronic structure, but this is properly
obtained only when the experimental measurements are combined with
quantum mechanical calculations of shielding. One could say that
measurement of the full shielding tensor (including its orientation in B 0 )
provides an order of magnitude more information than the isotropic value
alone.
For the full coupling contribution to the energy, this term must be summed
over all pairs of spins. For molecular solids, indirect coupling is generally
confined to intramolecular interactions, but it can extend across hydrogen
bonds between molecules and therefore it can be used to recognize the
existence of such bonding. Its relationship to connectivity is particularly
valuable for solid-state NMR.
where r is the distance between the two dipoles (see figure 2.3 ).
For a heteronuclear spin pair in the high magnetic field approximation, this
expression becomes (to first order in energy):
where θ is the angle between the internuclear vector and the magnetic field,
while D jk is the dipolar coupling constant between spins j and k. The latter
is given (in frequency units) by:4
Typical values of dipolar coupling constants are given in table 2.2 . The
tensor D has no antisymmetric component and, unless there is anisotropic
motion involving the two nuclear spin sites involved, it is also axially
symmetric. The isotropic component of D is zero, so dipolar coupling does
not affect resonance frequencies for solutions, although dipolar interactions
are a primary cause of relaxation.
Table 2.2. Some typical bond distances and corresponding dipolar coupling
constants
Table Missing
For polycrystalline powders, the distribution of orientations must be taken
into account, as shown above for shielding. In fact for a pair of isolated
heteronuclear spins (AX), the A spectrum will look like two superimposed
patterns of the type shown in figure 2.2(a) , one for each of the doublet
resonances, with reversal of direction for one of them. This is illustrated in
figure 2.4 . Such a powder pattern is known as a Pake Doublet ,5 the outer
limits of which correspond to θ = 0° and the intense “horns” to θ = 90° .
The separation of the horns is D , so the distance r AX can be obtained even
in the case of a powder. The X spectrum will be identical. The powder
pattern for an isolated pair of homonuclear spins will be a single similar
shape, but with a scaling factor of missing image file .
In fact, the classic case of a dipolar coupling is that of the proton spectrum
of gypsum, CaSO4 · 2H2 O, for which the protons in the water molecules
form relatively isolated spin pairs. Since this case involves homonuclear
interactions, the factor of missing image file , mentioned above, is
required for the spectrum. The unit cell contains two water molecules with
different orientations, so two doublets are generally observed for a single
crystal in a fixed setting of the crystal in B 0 , though for some special
settings only one doublet is observed. For one setting, a single line is
observed, showing that r HH for the two different water molecules can be
nearly at the magic angle θ = 54° 44′ to B 0 simultaneously. A full analysis
shows that the dipolar coupling constant is 30.7 kHz, so r HH is determined
to be 0.158 nm. Given the traditional difficulties of determining hydrogen
positions by X-ray diffraction, such NMR experiments are potentially
valuable.
For solids, the quadrupolar effect can be very important. Indeed, it often
dominates the appearance of spectra, since the quadrupole coupling
constant may be hundreds of megahertz. On the other hand, when the
nucleus in question is at a site of cubic symmetry, the EFG is in principle
zero. An example is the EFG at 33 S in a sulfate ion; however, it will only
be negligibly small if the ion retains its tetrahedral symmetry in the solid
state, whereas frequently the local environment lowers the symmetry.
If the quadrupolar energy is much smaller than the Zeeman energy, it can
be treated as a first-order perturbation (see chapter 6 for details). In such a
case, the energy, missing image file for nuclear spin I is given by:
The full quadrupolar energy term must be summed over all the types of
quadrupolar nuclei present. However, unlike the situation for dipolar
interactions (which occur for spins pairwise), the quadrupolar effect
involves single spins acting independently, which simplifies theoretical
treatments.
Since there are 2I + 1 energy levels (figure 2.7 shows the case for a spin-
missing image file nucleus), there will be 2I nondegenerate lines in the
spectrum for each value of θ other than 54° 44′.
Most quadrupolar nuclides have spin quantum numbers that are odd
multiples of missing image file (i.e., missing image file
missing image file , missing image file …). In such cases, the central
transition missing image file is unique because the first-order quadrupolar
effects on the two energy levels are equal. This is true even including cases
of nonzero asymmetry, since the quadrupolar energy depends on
missing image file , not m I (see equation 2.27). Therefore, the
quadrupolar contribution to the transition frequency is zero and is thus
independent of the orientation. The other NMR transitions (satellite
transitions ) are, however, strongly dependent on θ , so the intensity for a
polycrystalline sample will be spread over a wide frequency range. The
central transition resonance will thus stand out like a sore thumb. Figure 2.8
shows a powder pattern for an I = missing image file hypothetical case.
The spacing of the prominent horns (indicated by the arrow in figure 2.8 )
depends on the spin quantum number I ; in the I = missing image file
case, it is missing image file . The satellite transitions form subspectra as
indicated. Figure 2.9 provides an experimental example for an I =
missing image file system.
Figure 2.7. Energy levels for a quadrupolar nucleus. Left: zero quadrupole
coupling constant; right: schematic introduction of a small quadrupole
coupling constant (first-order effect).
However, when χ is large, as is often the case, the simple theory given
above is not adequate and second-order effects become prominent, as will
be described in chapter 6 . The case of nuclides with integral spin is also
different since there is no central transition (see inset 2.3). The only
nuclides in this category usually considered to be important are 2 H, 6 Li,
10 B, and 14 N. Of these, the first two do not generally give broad spectra,
since they have small quadrupole moments, and 11 B is usually preferred to
10 B, so only 14 N causes special difficulties.
Figure 2.10.
The average value of this function for isotropic motion is zero. Thus
molecular motion in solutions means that the anisotropies in σ and J do not
affect the resonance frequencies. Since the isotropic values of D and q are
zero, these interactions also have no significant influence on solution-state
resonance frequencies.
For a polycrystalline sample, the angle Θ can take all possible angles, just
as θ can. However, the angle β is fixed by the experimentalist. Therefore,
the factor missing image file acts to scale the powder pattern produced by
the range of values of θ . If β = 0 ° (rotation about the z axis), there is no
scaling. At the other extreme, β = 90 ° , the spread of the powder pattern is
reduced by a factor of 2, and if β = 54 ° 44′, the effects of anisotropies
vanish since the scaling factor is zero missing image file The width of the
signal is reduced as if by magic, so the term “ magic-angle spinning ”
(MAS) was devised to describe the phenomenon. The MAS technique is
now ubiquitous in solid-state NMR.
The right-hand side of figure 2.12 shows an example for 13 C. The spin
rate should be significantly greater than the static linewidth in order to
provide a fully narrowed single signal. Normally, spin rates in the region 3
− 30 kHz are used, though faster rates are now feasible. At the lower end,
these suffice to average 13 C shielding anisotropies, but higher speeds are
necessary to remove the effects of H,H dipolar interactions, and it is
impossible to completely average most quadrupolar effects.
2.8 RELAXATION
(c) Spin–lattice relaxation in the rotating frame. This describes the return to
equilibrium of transverse magnetization in the presence of an RF magnetic
field, B 1 , in the same direction. In this situation, the magnetization is said
to be spin-locked , because the relaxation time constant, T 1 ρ , is greatly
extended beyond transverse relaxation. It may be likened to T 1 but as
appropriate for a low magnetic field (B 1 ) rather than a high field (B 0 ).
Typically, values of T 1 ρ for 1 H in solids are 1− 20 ms (whereas for
mobile solutions they are equal to T 1 ). Values of T 1 ρ are of particular
importance in cross-polarization experiments (see section 3.4).
For solids, motions can be very complex and are often poorly understood.
In simple cases, for example molecular motion which is either isotropic or
involves rotation about a unique axis, T 1 and T 1 ρ pass through a well-
defined minimum as the motional rate increases, with a tendency to
increase toward infinity for very fast motions (as in mobile liquids) or very
slow motions (as for rigid solids). For most solids, mobility at the
molecular level is complex and the relaxation behavior as a function of
temperature is poorly understood. However, internal rotation about C–CH3
bonds tends to be relatively facile and so its effects can be well separated
from those of other motions. In such situations, relaxation times may show
several minima as a function of temperature. Figure 2.14 shows the case for
trimethylphosphine sulfide, which has two T 1 minima corresponding to
distinct motional processes, namely rotation of the methyl groups and
rotation of the whole molecule about the symmetry axis.
FURTHER READING
“High resolution NMR in the solid state” , E.O. Stejskal & J.D. Memory,
Oxford University Press (1994), ISBN 0 19 507380 0.
NOTES
2 Metal samples have so-called Knight shifts (which can be very large) in
their resonances. These will not be dealt with in this book. Paramagnetic
samples are also subject to special shift effects (see section 7.5).
4 Some texts define D jk include a negative sign in its definition. The key
point is that the energy expressed in equation 2.23 is lower when the
nuclear magnetic dipoles are parallel (if γ values for both nuclei are
positive).
5 The term “Pake Doublet” originally referred to the homonuclear case.
6 The symbol V is often used in the literature instead of eq for the EFG
tensor.
8 The symbol C Q is often used for this quantity, though χ is the IUPAC-
approved symbol.
3.1 INTRODUCTION
H He C
N F Si
P Fe Se
Y Rh Ag
Cd Sn Te
Xe Tm Yb
W Os Pt
Hg Tl Pb
Figure 3.2. The axis of a spinning body precesses around another axis (the motion of a gyroscope).
3.1
Figure 3.4. A pulse applied along x rotates the bulk magnetization vector about x .
The RF field strength associated with a pulse is a magnetic flux density and has
the unit tesla (T). The field associated with a pulse, B 1 , typically has a magnitude
of several milliteslas. If, say, this is 10 mT , a pulse duration, τ p , of 3 μ s would
produce a pulse angle of 90° for silicon-29 (from equation 3.1).
The strength of the resonant RF field is usually discussed in terms of an
equivalent rotation rate (the nutation rate of the affected spins):
3.2
Off-resonance
So far it has been assumed that the pulse has been applied on-resonance and
ν RF = ν NMR . In the general case, and the frequency of the
signal of interest, ν , depends on interactions in addition to the Zeeman one,
for example, shielding (see equation 2.18), so When
(it has to be so for a spectrum containing more than one
resonance), instead of the magnetization being rotated about B 1 , it rotates
about an effective field along a tilted axis , B eff , which is the resultant from B
1 and a residual static field B r = 2 π Δ ν /γ (B 0 is no longer effectively zero
in the rotating frame):
3.3
There are two main methods for exciting nuclei and generating spectra in
solid-state FT-NMR. The simplest experiment consists of a pulse followed
by the detection of an FID (an acquisition ). It is variously known as pulse-
acquire, single-pulse excitation (SPE), direct polarization, or direct
excitation. Direct excitation is the term used here. Alternatively,
magnetization can be transferred from an abundant spin such as 1 H to a
dilute one. This cross-polarization technique is described in section 3.4. The
previous section dealt with the excitation part of the direct-excitation
experiment. The other components and some of the parameters under the
control of the spectroscopist are introduced next.
The signal-to-noise ratio (S/N) is a useful measure of the quality of a spectrum and
the performance of the spectrometer. The noise is usually characterized by its
standard deviation and that, relative to the height of the signal in question, gives
the signal-to-noise ratio. Invariably, the spectroscopist seeks the highest possible
signal-to-noise ratio.
Figure 3.7. Two 31 P spectra acquired from the same sample but with different recycle delays. At the
shorter recycle, the broad, low-frequency shoulder on the narrow line is barely detectable and the
high-frequency signal is significantly underrepresented.
3.4
Figure 3.8. The amount of z magnetization at time t (relative to T 1 ) after a 90° pulse.
Figure 3.10. Spectra recorded with different recycle delays. The horizontal line is drawn at 72% of
full intensity. The maximum intensity is obtained with a 30 s recycle, although little signal is
sacrificed with a 10 s delay. In terms of the signal-to-noise ratio in any given total time, the optimum
recycle is around 3 s.
3.3.2 ACQUISITION TIME
The acquisition time is the time during which the signal (the FID) is
detected. Getting this right is important to the appearance, and information
content, of the spectrum. If the acquisition time is too long, the unnecessarily
acquired noise in the time domain results in additional noise in the spectrum.
If it is too short the FID is truncated . Fourier transformation of a truncated
signal results in “wiggles” at the base of the lines in the spectrum (figure
3.11 ).4 Although it is possible to compensate for a truncated FID with
additional mathematical processing (as is commonly done for two-
dimensional data sets) and an overlong acquisition can be multiplied by an
apodization function (see Further reading), it is far better to optimize the
acquisition time to start with, if the signal is sufficiently visible.
The receiver gain is the amount of amplification applied to the raw signal. The
receiver (the part of the spectrometer that digitizes the signal detected in the
NMR coil) has a finite range. Very-high-amplitude signals will overflow the
receiver and so will not be properly digitized. Significant clipping of the signal
leads to baseline artifacts in the spectrum (figure 3.12 ). Usually, the receiver
gain can be adjusted to avoid such problems, but in extreme cases (e.g., 1 H
observation) it may be necessary to add some attenuation to the signal
pathway. Too low a gain can result in artifacts when digitizing very weak
signals (digitization noise ).
missing image file
Figure 3.12. The “clipped” top and bottom of this free induction decay (FID) are characteristic of too
much signal prior to digitization. The subsequent Fourier transform results in the dips at the foot of the
line and the artifacts in the baseline.
3.3.5 DEAD-TIME
Signals extending over large spectral widths can be a problem in solid-state
NMR—that is, if the spectrum is to have a flat baseline.6 In any NMR
experiment, excitation by relatively high- voltage RF pulses (typically
hundreds of volts in solid-state NMR) is followed by the detection of a low-
voltage (typically microvolt) response. Ideally, acquisition needs to start
immediately after the pulse. However, if it is started too soon, the beginning
of the FID can be distorted by pulse breakthrough (real pulses are not perfectly
rectangular and the decaying “tail” of the pulse can be detected by the
receiver). The same spectrometer circuitry also has to cope with both high-
and low-voltage conditions and it tends to ring from the shock of the high-
voltage pulses. The time that it takes for the pulse to decay and the circuitry
to settle down is called the dead-time (see figure 3.14 ). No meaningful data
points can be obtained during the dead-time so it is not possible to follow the
evolution of the signal from its start. In practical terms, this results in a
distortion to the phasing of the spectrum (spectral phasing is determined by
the way the real and imaginary parts of a complex Fourier transformation are
combined, but that is something that is covered in other texts; see Further
reading). This is particularly noticeable when the spectrum extends over a
large frequency range as it does in the case shown in figure 3.15 . A first-
order phase correction can be applied to the spectrum to produce signals that
are in-phase, but that introduces a roll to the baseline. Matching the dead-time
to an integer number of dwell periods, together with numerical prediction of
the early part of the signal decay, can minimize the baseline roll.
Alternatively, the baseline correction routines in the spectrometer software
can be used to produce a flat baseline. Another technique for cases involving
large spinning sideband manifolds is to carry out the Fourier transform from
the top of the first rotary echo . Under magic-angle spinning (MAS),
orientation-dependent interactions give rise to rotary echoes. Echoes form
because the resonance frequency from a given crystallite changes as the
rotor position changes during spinning and each time the rotor returns to its
original position the resonance frequency returns to its original value. Rotary
echoes are illustrated in figure 3.15 (d). An echo can also be used (in fact it
is usual to do so, see inset 3.5) to record the broad lines from non-spinning
samples (the refocused signal at the top of an echo is a good approximation
to observing the signal from zero time).
Figure 3.15. The 199 Hg magic-angle spinning spectrum from Hg(SCN)2 acquired with a 500 kHz
spectral width and at a spin rate of 9.78 kHz. There was a 15 m s dead-time delay between the end of
the pulse and the first data point. (a) With no first-order phase correction. Note the flat baseline but
extensive distortion to the phases of the spinning sidebands. The centerband is the in-phase signal at -
1300 ppm. (b) The same data set but with a large first-order phase correction (- 3529°). The sidebands
now all have the same phase but the baseline rolls. (c) The same data set but transformed from the top
of the first rotary echo shown in (d). Only a small (96°) first-order phase correction was necessary to
produce this spectrum.
A significant amount of signal can be lost in the spectrometer dead-time for the
broad lines encountered with non-spinning samples. When shielding anisotropy is
negligible most of the missing signal can be recovered using a solid (or
quadrupolar) echo pulse sequence (see also section 4.4.2):
missing image file
1
The H spectra on the right were obtained (a) with and (b) without a solid echo.
The integrated intensity for the band in (b) is 23% of that shown in (a) from the
same sample. In (b) the dead-time between the pulse and the first point of the FID
was 10 μ s.
missing image file
3.3.7 DECOUPLING
It is sometimes necessary to use a very long acquisition time (> 100 ms)
to avoid truncating an FID. This situation may occur for samples with high
molecular mobility such as gels or rubbers. As these also tend to have short
relaxation times (allowing short recycle delays), it is easy to reach the duty-
cycle limit. However, such samples also tend not to require very high
decoupling power (the molecular motion partially averages the dipolar
interactions involving 1 H) so the duty-cycle limit can be exceeded providing
the decoupling power is reduced.
For organic materials, it is well worth using as much decoupling power
as is available, up to the limit the probe can withstand, as figure 3.16
illustrates. However, setting too high a decoupling power can result in a
spectacular electrical breakdown in the probe (known as arcing ). Minor
arcing can result in an increased and nonuniform noise pattern in the
spectrum. Serious arcing can obliterate the signal that is being detected (see
figure 3.17 ).
Table 3.1. Common chemical shift reference compounds (including some for
quadrupolar nuclei) relative to recognized standard compounds (see the reference in
footnote f ).
Nuclide Reference for solid-state NMR (chemical shift / ppm)
1
H Adamantane (1.9) a
Glycine (8.5) b ,c
13
C Adamantane (38.5, CH 2 ) c ,d
Glycine (176.5, COO – ) c
15
N NH 4 15 NO 3 (–5.1, NO 3 ) c
Glycine (–346.8)
19
F C 6 F 6 ( –164.9) e
29 tetrakis(trimethylsilyl)silane (–9.8,–135.4)
Si
31 CaHPO 4 · 2H 2 O (1.0)
P
77 (NH 4 ) 2 SeO 4 (1040)
Se
119 Sn(C 6 H 12 ) (–97.4)
Sn
199 [Hg(dmso) 6 ][O 3 SCF 3 ] 2 (–2313)
Hg
7
Li LiCl (0) f
11 BF 3 /O(CH 2 CH 3 ) 2 (0)
B
17 H 2 O (0)
O
23
Na NaCl (0) f
27
Al Al(NO 3 ) 3 (0) f
45
Sc Sc(NO 3 ) 3 (0) g
51 b -NaVO 3 (–519) (solid)
ν
59
Co K 3 Co(CN) 6 (0) f
a
At high (>10 kHz) spin rates (when the centerband is well separated from the
spinning sidebands).
b
When used under high-resolution, multipulse conditions. Literature values vary
considerably so this value should be used with caution.
c
This is the highest frequency centerband in the spectrum.
d
Ξ = 25.145 970 and 25.145 743% for the CH2 and CH, respectively (see text).
e
There is some debate about this value. cfc l3 is no longer freely available so
resolving this issue is a problem.
f
In aqueous solution; details given in Pure Appl. Chem. 80 (2008) 59.
g
0.11 M in 0.05 M HCl.
Referencing methods that use documented absolute frequencies for
different nuclei, defined through the parameter Ξ , relative to the 1 H
resonance from tetramethylsilane in dilute solution in chloroform (Ξ = 100%
)8 can be used in the solid state, although this may require some ingenuity in
the absence of the deuterium lock used to do this in solution-state NMR.
(With the possible exception of biochemical solid-state applications, solid-
state experiment times are generally too short and lines are too broad for any
drift in magnetic field to be noticed, so a deuterium lock is not usually used
in solid-state measurements.)
For solid-state NMR it can be used either neat (if it can be contained) or on a
solid support. It is useful for calibrating temperature as a function of spin
rate (see figure 3.18 ) under “ambient” conditions and for giving an absolute
temperature. Because most probes can operate over a wide temperature
range, other temperature-sensitive substances have to be used for
temperatures at which methanol is not useful (below − 90 °C or above +60
°C). One such is solid lead nitrate (see figure 3.18 ). The chemical shift of
the 207 Pb signal is very sensitive to temperature (0.70 ppm / °C) and can be
used to determine the temperature difference from a known point (there is no
accepted absolute relationship between temperature and the chemical shift of
the single lead resonance), for temperatures up to about 250 °C (it
decomposes at 290 °C). The shape of the line (see figure 3.19 ) also gives an
indication of the temperature variation in the sample (this increases with
extremes of temperature and can be several degrees depending on the shape
and size of the sample space). The only major drawback with using lead
nitrate is its toxicity.
Figure 3.19. 207 Pb spectrum from lead nitrate at nominal temperatures 25 and 150 °C. In the high-
temperature spectrum, the parts per million spread of the signal equates to a temperature range from
around 133 to 158 °C.
The impact that these two factors might have on experiment time is
exemplified in inset 3.6 and illustrated in figure 3.20 .
The right conditions for polarization transfer occur when the 1 H and X
RF fields fulfill the Hartmann–Hahn match condition (the fields are then said to be
matched ):
missing image file or, from equation 3.2, missing image file
3.7
constant). It is usually safer if the former is varied while the contact pulse is
kept constant as this is less likely to result in too much RF power being
applied to the probe, since missing image file at the matching
condition (inset 3.7 relates to the measurement of RF power). Figure 3.22
shows a typical 13 C match profile from hexamethylbenzene (HMB).9 At a
modest spin rate (3 kHz), a well-define d maximum is observed (this is
relatively uncommon behavior—the match profile is broader and flat-topped
for more typical materials).
missing image file or missing image file , where P is the power in watts and V
pp is the peak-to-peak voltage (at an impedance of 50 W ).
A decibel value is the ratio of two powers: P(dB) = 10log10 (P1 /P2 ) When
quoted as dBm P 2 has a reference value of 1 mW.
So 1 mW ≡ 632 mV (0 dBm), 100 W ≡ 200 V (50 dBm), and 1 kW ≡ 632 V (60
dBm).
For HMB the match behavior becomes more complicated at higher spin
rates (see section 5.1.3) and in other materials similar complexity can arise at
low spin rates when hydrogen is dilute in the sample or when there is a high
degree of molecular motion (as there is for adamantane, which is commonly
used in setting up experimental conditions for a spectrometer).
Any X pulse, including a contact pulse, will result in a signal being
generated by direct excitation as well as, and independent of, that originating
from cross polarization. This is most noticeable for nuclei with high natural
abundance such as 31 P or for isotopically labeled samples where the direct-
excitation signal has a high intensity. This unwanted signal is suppressed
using a phase cycle .
Phase Cycling
The phase cycle is a key part of most solid-state NMR experiments. As has
already been discussed, the pulses in a pulse sequence are labeled with a
phase. The receiver has a phase too (see inset 3. 8). Over a cycle of
successive repetitions of the pulse sequence, it is usually the case that these
phases change—the phase cycle. In the simplest of cases, the phase cycling
uses the quadrature phases x , y , – x , and – y (or 0, 90, 180, 270) to
compensate for any missetting of the phases (exactly how this works is not
imp ortant here). The receiver phase simply needs to follow the pulse. So the
pulse-acquire experiment might be phase cycled as:
The “ raw ” NMR signal is split into two to give “ real ” and “ imaginary ”
components that have a phase difference of 90°. This is a necessary precursor to
the complex Fourier transform that converts the time-domain signal into the
frequency spectrum. The details of the way the signal is split are not important
here–and in any case tend to change with developments in the electronics within
the spectrometer. The important point is that a receiver nominally with phase x will
detect both the x and y components of the signal.
Table Missing
Table Missing
Over the four repetitions of the pulse sequence in this phase cycle, the cross-
polarization signal is always detected as positive (relative to the receiver
phase) but the sum of the direct-excitation signal is zero over the phase
cycle. The alternating reversal of the phase of the 90° pulse is sometimes
referred to as spin-temperature inversion .
Designing complex phase cycles (like the one illustrated in inset 3.9) is
an advanced skill and takes some practice (and understand). An error in the
phase cycle is a common cause of unexpected behavior from a new pulse
sequence. See the texts in the Further reading for detailed information on
devising phase cycles.
During the contact time, the X magnetization builds up toward a steady state
corresponding to an equilibrium between the 1 H and X magnetizations. At
the same time the spin-locked 1 H and X magnetizations decay through spin–
lattice relaxation in the rotating frame (section 2.8) at rates of
missing image file a nd missing image file respectively. These processes
are summarized in figure 3.23 . If missing image file then
missing image file
The phase cycle is also a key part of the coherence selection in multiple quantum
experiments. Here the quadrature phases are not the only ones that can be used; 45,
90, 135 … are found in experiments where double quantum coherences are
selected and 30, 60, 90 … can be used for triple quantum coherence selection.
Figure 6.16(a) shows, a two-pulse, triple-quantum/single-quantum correlation
experiment that has the following 24-repetition phase cycle:
Table Missing
Table Missing
As this case illustrates, phase cycles can be many repetitions long. So that the
phase cycle can do the job it is designed for, it is always a good idea to acquire an
integer multiple of the number of repetitions in a phase cycle. That may not be as
trivial as it sounds for a complex two-dimensional experiment with a many-
repetition phase cycle—particularly if a long recycle delay is needed and the total
time required becomes prohibitive.
can be safely ignored and the amount of signal (S ) obtained with a contact
time t c is given by equation 3.8
Figure 3.25. (a) The flip-back modification to the cross-polarization pulse sequence and (b) 13 C
spectra from mannitol obtained with a 60 s recycle delay with (on the left) and without (right) flip-
back (all other acquisition conditions being equal). Without flip-back, a 300 s recycle delay is required
to produce the equivalent result—increasing the experiment time by a factor of five.
It was noted at the beginning of this chapter that 1 H and, to some extent, 19
F are special cases. This is because, despite the most advanced experimental
techniques available at the time of writing, 1 H and 19 F linewidths, for true
solids, remain stubbornly high (often several 100 Hz) and this fact limits the
chemical information available from a spectrum. This is in stark contrast to
the solution state, where a proton NMR spectrum is usually the starting point
for the characterization for any soluble, proton-containing material.
Solution-state linewidths are typically of the order of 1 Hz. In the solid state,
strong homonuclear dipolar coupling is the main contributor to high
linewidths. For 1 H, the problem is compounded by its small chemical shift
range (~20 ppm) and that results in poorly resolved spectra (see inset 3.10).
The chemical information from a 1 H spectrum may be limited to the general
identification of signals as aliphatic, aromatic or, when present, strongly
hydrogen bonded (which usually have a distinct high-frequency shift). The
situation is slightly better for 19 F because the chemical shift range is over
400 ppm and useful chemical information can be usually obtained. Examples
of solid-state 1 H and 19 F spectra are given in figure 3.26 .
Spectral resolution is a measure of the degree to which two closely spaced signals
in a spectrum can be separately identified. The resolution depends on linewidth
and line separation as illustrated in the figure (right), which shows two lines with
Gaussian shape and half-height width Δ ν , (a) at the same chemical shift, (b) 0.8Δ
ν apart—unresolved (the flat top is characteristic of this situation), (c) Δ ν apart—
just resolved (the limit of resolution), and (d) 2Δ ν apart—resolved.
Figure 3.26. Examples of solid-state 1 H and 19 F spectra. (a) Naturally resolved 1 H spectrum from
the mineral octosilicate (spin rate 5 kHz). (b) 1 H spectra from 3-methoxybenzoic acid, magic-angle
spinning (MAS) at 15 kHz (upper trace) and with homonuclear decoupling (a windowed phase-
modulated Lee-Goldburg, wPMLG, spectrum) and spinning at 10 kHz. (c) 1 H spectrum from an
organic compound in a silica matrix recorded at a spin rate of 60 kHz at a field of 21.1 T. (d) 1 H high-
resolution (HR) MAS spectrum from an organic compound bound to a swollen polystyrene bead. (e)
19 F spectrum from polyvinylidenedifluoride. (f) 19 F spectrum from octafluoronaphthalene . Note the
different scale ranges. Figure 3.26 (c) supplied courtesy of Dr Anne Lesage.
FURTHER READING
“Understanding NMR spectroscopy” , J. Keeler, John Wiley & Sons (2005), ISBN 978 0 470 01787 6.
PULSE NMR AND PRACTICAL ISSUES RELATING TO SIGNAL
PROCESSING
“Modern NMR techniques for chemistry research” , A.E. Derome, Pergamon Press (1987), ISBN 0 08
032513 0.
“High-resolution NMR techniques in organic chemistry” , T.D.W. Claridge, Elsevier (2009), ISBN
978 0 08 054818 0.
NOTES
1 The axes of the rotating frame should be labeled differently to those of the fixed laboratory frame
(say, x ′ , y ′ , and z ′ instead of x , y, and z ). However, it is customary to omit the distinction. Phases in
this sense are relative labels and have no absolute physical meaning.
2 This rotation is what is expected from the classical physics of two interacting orthogonal magnetic
fields (the bulk magnetization and B 1 ; in the rotating frame and on-resonance, B 0 disappears).
Conventions for the direction of rotation vary—consistency is the important factor in describing an
experiment.
3 There is also a trade-off between pulse angle and recycle delay. If T is known, then the pulse angle,
1
θ , that results in the maximum amount of magnetization in the xy plane for a chosen recycle delay is
given by cosθ = exp(-recycle delay /T 1 ). This angle is known as the Ernst angle.
4 These sinc wiggles appear only if the signal is zero-filled (as is usually the case, see inset 3.10) prior
to Fourier transformation.
5 In fact, modern spectrometers automatically oversample (which is equivalent to using a very high
spectral width) and then digitally filter and downsample to give a more manageable number of data
points (all usually hidden from the spectrometer operator). This process is unlikely to result in folded
back signals.
6 Sometimes this may simply be a case of aesthetics, but if intensity information is required a flat
baseline is usually essential.
7 “Continuous wave” is a term in widespread use, but it should be noted that for solid-state NMR the
decoupling is not actually continuous. Here, “continuous wave” is merely a label for decoupling with
no amplitude or phase modulation.
8 For details see Pure Appl. Chem. 80 (2008) 59.
9 Hexamethylbenzene (HMB) is often used as a set-up compound for 13 C because it gives a well-
defined match profile irrespective of other instrument variables such as decoupling efficiency or
shimming. Adamantane (see inset 7.8) is a highly mobile solid that gives very narrow lines and is a
good check for lineshape.
10 In figure 3.21 the spin-lock and decoupling pulses have the same phase so the spin-lock is
maintained even when there is an increase in RF field for decoupling. The spin-lock is lost if the 1 H
irradiation is turned off, as it would be for a dipolar dephasing experiment (see section 7.2.4), or if
more complex decoupling schemes than simple CW irradiation are used.
11 1 H,1 H coupling enables spin diffusion to be efficient among the protons and this leads to
increased polarization transfer, making it more likely that a signal enhancement closer to the
theoretical missing image file maximum will be achieved.
12 When missing image file 1 H decoupling has no detectable impact on the chemical shift of the
observed X nuclei. If this condition does not hold (as it might not for 19 F observation with 1 H
decoupling), then a small shift in the X resonance frequency may be observed. This is known as the
Bloch–Siegert effect.
13 Because these materials contain no 1 H, they are also ideal for use when observing 13 C using 1 H
to 13 C cross polariz ation (in effect they are invisible). They will, however, contribute signal to a
direct-excitation spectrum.
CHAPTER 4
4.1 INTRODUCTION
The previous chapters have deliberately avoided delving deeply into the
theory underlying solid-state NMR. Indeed many solid-state NMR
experiments can be understood in terms of the vector model familiar from
the solution state (and discussed in section 3.2), so that most experiments
can be interpreted using straightforward qualitative pictures of the different
NMR interactions and their effects on the spectrum. As the subject has
progressed, however, the complexity of experiments and the associated
theoretical background have considerably increased. This chapter aims to
provide an introduction to the various theoretical tools used in solid-state
NMR. The Further reading lists some more detailed and thorough texts, but
this concise overview should give the reader sufficient insight to be able to
explore further and to tackle the NMR literature with confidence. Those
only interested in using the results of solid-state NMR will probably prefer
to skip over this chapter and go to the applications of solid-state NMR,
ignoring the references to theory when some of the more complex
experiments are discussed.
So while the full J-coupling Hamiltonian involves both the spin coordinates
and the entire electronic wavefunction, the corresponding spin Hamiltonian
just involves the states of the two relevant nuclear spins and a coupling
constant, J IS , which results from integrating out the other degrees of
freedom.
where the magnetic field, B 0 , has its conventional alignment along the z
axis, and Î z is the operator for the z component of the nuclear spin angular
momentum. The Schrödinger equation is trivial to solve for this
missing image file since the 2I + 1 spin states are already eigenfunctions
of Î z
missing image file 4.2
This recovers the simple expression for the energy of a nuclear spin in a
magnetic field presented in chapter 2 :
These names are derived from their effect on the Zeeman eigenstates:
Note that the factor of missing image file has been dropped from the spin
operators. This corresponds to the normal NMR convention of expressing
the spin Hamiltonian and its components (such as coupling constants) in
terms of frequency rather than energy.
Because the other interactions are usually small in comparison with the
Zeeman interaction, the eigenbasis of the total nuclear spin Hamiltonian is
the same as the Zeeman eigenbasis figure missing (except when large
quadrupole interactions are present, as discussed in chapter 6 ). Moreover,
components of the interaction Hamiltonian that do not commute with the
Zeeman interaction can be neglected .
Matrix trace The trace of a matrix is the sum of its diagonal elements
Handy relation : ( AB ) −1 = B −1 A −1
Transpose and conjugate transpose Transposing a matrix A (common
symbol A T ) involves swapping matrix elements across the diagonal. The
conjugate transpose , A † , involves taking the conjugate of the matrix
before or after transposition:
Handy relation : ( AB ) T = B T A T
The final form of the equation 4.18 depends on the nature of the
interaction. For instance, the rank-2 component of the J (or indirect dipole–
dipole) coupling is either eliminated by molecular tumbling in the solution
state or can be merged into the much larger direct dipole–dipole interaction
(see equation 8.3). Hence the J coupling between two spins I and S is
normally reduced to its isotropic component
where the scalar (dot) product between the spin operators expands to
figure missing In the homonuclear case, I and S have the same Larmor
frequency, ν I , and the Zeeman Hamiltonian is
In this case, only the figure missing terms of figure missing commutes
with figure missing and the coupling Hamiltonian can be reduced to
which is equivalent to the expression for the nuclear spin energies given in
equation 2.23.
The final term required for this quantum treatment of the NMR experiment
is the Hamiltonian for the radiation used to excite the nuclear spins. The
Hamiltonian for a nuclear spin interacting with electromagnetic radiation is
where B 1 is the intensity of the oscillating magnetic field and ø is its phase
(with zero conventionally corresponding to initial x phase). ν RF denotes
the frequency of the radiation from the radiofrequency (RF) transmitter.4
The reason for the leading factor of 2 will appear later.
that is, the xy axes rotate around the z axis with angular frequency ω RF
(the use of angular frequencies avoids factors of 2π).
that is, one component of the RF is now independent of time in the new
(rotating) frame, while the other is rotating at an angular frequency of 2ω
RF .
The transformation into the rotating frame also affects the Zeeman /
chemical shift Hamiltonian . The Zeeman + shift Hamiltonian in the
laboratory frame is
(for a single spin), where ν 0 refers to the Larmor frequency (including the
shielding term). The effect of this Hamiltonian is to cause the net nuclear
magnetization to precess about z at its NMR frequency. The effective
precession rate in the rotating frame will be missing image file Hence the
effective Hamiltonian in the rotating frame is
The matrix elements, figure missing are identical between the different
systems, hence the measured expectation value reduces to
missing image file 4.39
that is, the only quantities we need to define are the ensemble-averaged
coefficients missing image file Hence it is sufficient to define the state of
the system using the matrix, ρ , with elements
where tr is the matrix trace (see inset 4.2). We lose information about the
wavefunction of an individual system, but this is not a great loss as the
NMR experiment invariably involves measurements on large ensembles.
For example, the equilibrium state of the nuclear spins within a magnetic
field involves net magnetization along the z axis, with net excess of spin
magnetization, M 0 /2, in, say, the α state. The density matrix6 for the
equilibrium state would therefore be
We can then obtain expectation values for the different angular momentum
operators of equation 4.9, for example:
Note how the diagonal elements of the density matrix, figure missing ,
directly correspond to the populations of the states. The physical
significance of the off-diagonal elements is less immediately obvious, but
consider the density matrix
missing image file 4.45
Clearly the average populations of the α and β states are zero, but the
system is still in a well-defined state. This is impossible classically since a
two-level system must be in one of the two available states at any one time.
Quantum mechanics, however, allows different states of existence,
coherences , which involve mixtures of the eigenstates. Hence we need a
matrix to hold the populations of the states (diagonal elements) and the
coherences (off-diagonal elements) that can be created between the
different states.
Figure 4.1 illustrates different ways of representing the state of the set of
nuclear spins. The vector model, figure 4.1(a) , is useful when there is a
single spin type or when couplings between spins can be ignored. It is
difficult, however, to represent coherences other than simple x , y , and z
magnetization in this way. Especially when problems involve an isolated
pair of spins, it can be useful to consider an energy level diagram such as
figure 4.1(b) . The relative populations of the four possible states are
indicated by the occupancy of each level while the arrows mark possible
coherences between the states. Coherences can be illustrated with such
diagrams, for example, the single-headed arrow would indicate a coherence
(of order + 2) between the figure missing and figure missing states, but
this could be confused with a transition or a relationship between the
populations of the states involved. The density matrix representation, figure
4.1(c) , provides the most compact representation of the different possible
states of the system.
The NMR experiment can now be described in this formalism, starting with
the density operator that describes the initial equilibrium state of the
system. At equilibrium there are no net coherences between the states, and
the relative populations of the eigenstates are given by the Boltzmann
distribution. The equilibrium density operator is7
where missing image file is the total nuclear spin Hamiltonian and Z is a
normalization factor (the partition function). Since the nuclear spin energies
are so small in comparison with thermal energies, the exponential can be
expanded
where Z equals the number of spin states, 2I + 1, in this limit (see inset 4.3
for an explanation of how to take exponentials of matrices such as
missing image file ). missing image file is the identity operator and so
represents an equal distribution of population between the different states.
This term is of no interest (since it does not evolve under
missing image file ) and can be dropped. The second term corresponds to
the tiny changes to the populations caused by the lifting of the degeneracy
of nuclear spin states by the spin Hamiltonian missing image file .
The density operator approach comes in to its own when multiple spins are
involved, but it is useful to start with some simple experiments that can be
easily visualized in vector model terms.
The system Hamiltonian (in a frame rotating at the same rate as the RF)
consists of a small chemical shift offset term
Because missing image file and missing image file commute (they are
simply proportional to each other), the Liouville–von Neumann equation,
equation 4.50, reassuringly predicts no evolution of the density operator
missing image file
Handy relation : if two matrices commute, then the same matrix V will
diagonalize both matrices, that is, they share a common eigenbasis.
where k is a scalar.
where the elements of the exponential of the diagonal matrix Λ are simply
[exp( Λ )] ii = exp( Λ i ).
that is, missing image file after a missing image file pulse. This
corresponds to the nuclear spin magnetization now being aligned with the y
axis.
When the RF is turned off, the Hamiltonian returns to equation 4.59 and the
propagator for the following period of free precession (evolution under the
system Hamiltonian in the absence of RF) is
This simply corresponds to rotation about the z axis at rate Δ. Hence the
density operator is
Note that evaluating the NMR signal as tr( missing image file Î + )
corresponds to detecting p = − 1 coherences of the density matrix. For
example, if the density matrix is a hypothetical pure − 1 coherence,
missing image file = MÎ − , the observable NMR signal at this point is
The comparison between the vector model and density operator treatments
of this problem is illustrated in figure 4.2 .
Echoes of various kinds form the central building blocks of many NMR
experiments. Two of the most important of these, the simple (or Hahn )
echo and the solid (or quadrupolar ) echo , are illustrated in figure 4.3 .
The simple spin echo is used to “refocus” the effect of Hamiltonian terms
that are “linear” in the spin of interest, such as the chemical shift and
heteronuclear couplings. After initial excitation, in this case using a
missing image file pulse to create y magnetization from the equilibrium
state, the xy coherences (or equivalently + 1 and − 1 coherences) evolve
during the free precession period, τ . Including just linear terms, the free
precession Hamiltonian might be:
where Δ is the offset from the origin of the rotating frame, and J is a scalar
coupling between I and S spins. The problem can, however, just be
expressed in terms of the I spin, since the sub-spectra corresponding to
different states of the S spin are fully independent. For example, if S is a
spin- missing image file , then we can consider an effective offset, Δ′ = Δ
± J /2, reducing the Hamiltonian to missing image file
at the end of the first τ period and before the inversion pulse.
The missing image file pulse will invert the sign of the y magnetization
and leave the x component unchanged, that is, the density operator after the
pulse is:
Figure 4.3(c) shows the coherence pathway diagram that applies to the
spin-echo (and also the quadrupole-echo) experiment. These diagrams
summarize the changing state of the spin system during the experiment in
terms of a “pathway” showing the order of coherences (zero quantum,
single quantum, etc.) present at different steps of the experiment. The
pathway always starts at zero order, corresponding to equilibrium z
magnetization, and finishes with observable p = –1 coherence
magnetization during the acquisition period.
During free precession periods (no pulses), the different coherences evolve
but do not change order. For instance, initial Î x magnetization will evolve
under the influence of a heteronuclear coupling to a spin S to create mixed
coherences of the form figure missing These are still, however, single
quantum coherences, and so the coherence pathways are horizontal during
such delay periods. By contrast, RF pulses mix coherences between the
different orders that the spin system can support, allowing the pathway to
jump between coherence orders. It is important to note that the coherence
pathway only summarizes the (desired) coherences present at a particular
point and does not predict the detailed state of the system. For instance, the
presence of an echo at 2τ must be deduced from analytical treatment of the
density operator.
Figure 4.3(b) shows the pulse sequence for the solid (or quadrupolar) echo
experiment. This has strong similarities to the simple spin echo (indeed the
coherence pathway diagram is identical), but the details are different since
the goal is to refocus the evolution under “bilinear” Hamiltonians, such as
the quadrupole or homonuclear coupling Hamiltonians.10 In particular, the
refocusing pulse is a 90° pulse; a 180° pulse has no effect on the evolution
under bilinear Hamiltonians, see section 4.7.3.
This may seem a great deal of effort to describe the basic NMR experiment,
and, indeed, the density operator treatment for a single spin-
missing image file is equivalent to the much simpler description of the
NMR experiment provided by the vector model of NMR. The density
operator approach comes into its own, however, for treating systems with
more than two levels.
It was noted without detailed explanation in section 2.5 that the splittings
due to homonuclear couplings were 50% larger than corresponding
heteronuclear couplings. This first example shows how this factor arises.
The matrix representation for terms such as figure missing is simply the
product of the matrices12 for Î + and figure missing Alternatively, the
matrix elements can be determined directly by considering the effect of the
individual operators on the appropriate spins. For example,
figure missing involves applying the lowering operator to S in the α state
and the raising operator to the I spin in the β state
Note the block structure of all the matrices, where the blocks correspond to
the different values of m I + m S , that is, the eigenvalues of the sum z
operator, missing image file The Hamiltonian matrix of equation 4.78, for
example, has non-zero elements in a 1 × 1 block corresponding to
missing image file = 1, a 2 × 2 block corresponding to the two states with
missing image file and missing image file and a final 1 × 1 block
corresponding to missing image file This block structure also applies to
the density matrix, with a block linking states with missing image file and
missing image file corresponding to coherences of order p = k – l . Hence
the system of two spin - missing image file nuclei, as previously
illustrated in figure 4.1 (c), can support coherences of order 0 (the
populations of the different eigenstates plus zero-quantum coherences
between, e.g., missing image file and missing image file detectable
NMR coherences of order ±1, plus multiple-quantum coherences of order
±2 between the αα and ββ states.
where Û is the propagator over the dwell time, Û ≡ Û (Δ t , 0). Since the
propagator for each dwell time is the same, we can write Û ( n Δ t , 0) = Û
(Δ t , 0) n . Equation 4.84 is most easily evaluated in the eigenbasis of the
Hamiltonian. The matrix of equation 4.78 is readily diagonalized thanks to
its block structure—only the central 2 × 2 block will have nontrivial
eigenvalues and eigenvectors:
Transforming missing image file and missing image file into the new
eigenbasis gives
Expanding out the matrix products of equation 4.84 gives (another exercise
for the reader):
If we create a new matrix with elements missing image file we can write
where ν I and ν S are the nutation frequencies due to the RF on the I and S
spins respectively. For simplicity we assume that the RF is on-resonance
for both spins. The Hamiltonian matrix in the product Zeeman basis is
Note how the ν I elements connect states in which the first (I ) spin flips,
while the ν S elements correspond to allowed NMR transitions of the S
spins.
Despite its elegance, the eigenvalues and vectors of this matrix are not
obvious, and there are large terms (those involving the RF) that are off the
diagonal. The situation is improved using a basis that puts the dominant
terms along the diagonal by rotating the axis system so that the new z axis
is along the RF spin-lock axis. The required 90° rotation about y simply
permutes the labels on the operators (x → z → − x → − z → x ) to give
Since the RF field strength must significantly exceed the dipolar coupling
for a good spin-lock, figure missing the secular approximation can be
invoked to drop the off-diagonal terms connecting different diagonal
elements, leading to
the mixing of the central states can be seen to interconvert figure missing
and figure missing It is left as an exercise to show, using the approach set
out in section 4.5.1, that the evolution of the density operator in the
conventional basis will be
missing image file 4.100
where missing image file is the sum magnetization operator. The second
term is a flip-flop term interchanging states figure missing and
figure missing while the third interchanges figure missing and
figure missing . If the spin-lock is strong, that is, figure missing then
the secular approximation can be invoked to discard terms that do not
commute with ν HH missing image file The Hamiltonian then reduces to
The initial density operator (in the tilted frame) is missing image file This
commutes with the RF Hamiltonian (i.e., the magnetization is spin-locked).
Hence the density operator will evolve simply under the influence of d
missing image file . As discussed above, this flip-flop term has the effect
of interconverting I and S spin magnetization.
Figure 4.4. Definition of the Euler angles (Rose convention) * for the
transformation of the original axes (x , y , z ) to a new axis system (x ′ , y ′ ,
z ′ ). The axes are first rotated by α about the z axis to bring the y axis onto
the nodal line defined by xy and x ′ y ′ planes. This is followed by rotation
about the new y axis through β and then a final rotation through γ is applied
about the final z axis (z ′ ).
Figure 4.4 shows the most common definition of the three Euler angles, α ,
β, and γ , specifying the rotation of one coordinate system to another.
Unfortunately, the definition of the angles is not unique and different
systems have been used in the literature. Care is needed if reporting Euler
angles or when compiling orientational information from different sources.
The tensor in the new basis is given by A ' = RAR -1 where the rotation
matrix R is
The final matrix A' is not shown as it is rather cumbersome and not
particularly informative. The main text discusses an alternative
representation of tensors that is better adapted to frame transformations.
where R iso is the isotropic component, and ζ and η are the anisotropy and
asymmetry respectively (see inset 2.2). Note how the spherical tensor
representation captures the rotational symmetry: the rank-0 isotropic term
is independent of orientation, while the anisotropy is contained within the
rank-2 terms.
At high field, the only components that survive the secular approximation
are the entity image missing and entity image missing components of
the tensor in the laboratory frame. The isotropic term is, by definition,
independent of the axis system, and so the problem of finding the high-field
Hamiltonian involves determining entity image missing after applying a
series of transformations to a tensor starting in its principal axis system.
where missing image file are elements of the rank-2 Wigner rotation
matrix . These are given by
missing image file 4.108
where the “reduced” matrix elements missing image file (β ) are solely
functions of the β angle.15
where the rotor position, α , is time dependent ( α 0 and ω r are the initial
rotor angle and spinning frequency respectively). The β angle corresponds
to the angle of the rotor with respect to the magnetic field, that is, the magic
angle θ m . The γ angle can be arbitrarily set to zero, as discussed above.
Unless the spinning rate is much larger than the interaction (in which case
the oscillatory terms corresponding to n = ±1 and n = ±2 are negligibly
small), the n = ±1 and n = ±2 in equation 4.110 will lead to some
oscillatory evolution of the density matrix. In terms of the NMR signal, this
corresponds to the formation of rotary echoes , in which the signal evolves
during the rotor period, but is “refocused” to create a rotary echo at each
complete rotor period (as required by equation 4.112). This series of
echoes, with a period equal to the rotor period, corresponds to a set of sharp
peaks separated by the spinning frequency in the spectrum, that is, a
manifold of spinning sidebands. See figure 3.15 for a practical example of
rotary echoes.
AHT applied to sample rotation has already been discussed in section 4.6.
In this simple case, it was sufficient to average over the time dependence to
obtain the average Hamiltonian to first order. By contrast, when dealing
with strong time-dependent RF, it is generally necessary to transform the
Hamiltonian into an interaction frame that allows the average Hamiltonian
to be readily computed. The total Hamiltonian is first separated into a term,
figure missing , containing the NMR interactions, etc., and a time-
dependent component, figure missing for the RF:
4.115
This may seem obscure, but the rotating frame of reference provides a good
example of an interaction frame; U RF (t ) in this case is exp(iω NMR tÎ z
), representing rotation about the z axis at the Larmor frequency, ωNMR ,
and missing image file is now the familiar rotating frame Hamiltonian.
we can write down the Hamiltonians for the successive steps of the cycle.
The transformation into the interaction frame for the period between the
first two pulses corresponds to relabeling the axes using y → z , z → –y ,
and so the interaction-frame dipolar Hamiltonian during this period is
The average Hamiltonian over the entire period is (to first order):
Since the expression for missing image file in equation 4.121 is linear in
missing image file we can compute the first-order average Hamiltonian
for different interactions independently, that is, missing image file Hence
equation 4.128 will be true for an arbitrary collection of homonuclear
coupled spins. This does not apply to higher-order contributions to the
average Hamiltonian, and calculating these higher-order “cross-terms”
between different components of the Hamiltonian is one of the major
challenges to applying average Hamiltonian approaches beyond first order.
where T is the period (e.g., the rotation period in the case of MAS, or the
sequence period in the case of periodic RF) and m is the index of the
Fourier series. However, since most functions can be expanded as a
convergent Fourier series, this is not a major restriction. Moreover, more
than one time dependence (e.g., magic-angle spinning plus time-dependent
RF) can be accommodated by expanding with respect to more than one
Fourier index.
The density operator, NMR signal, etc., can then all be expressed in terms
of Fourier series; for example, the NMR signal for MAS problems becomes
Without going into the mathematical details, Floquet theory provides the
machinery to calculate the A m,j coefficients, starting from the Fourier
series for the Hamiltonian (plus the initial density matrix). Note how
Floquet theory imposes no restriction on the sampling of the NMR signal
(equation 4.130 is valid for all t ) and naturally distinguishes centerbands
and sidebands.
As part of deriving the interaction frame for the WHH-4 sequence above,
the effect of the RF pulses was described in terms of “rotations” of the spin
operator components of the Hamiltonian. While the 90° rotations of WHH-
4 can be readily expressed in terms of Cartesian axes, working with
arbitrary rotations in this framework becomes cumbersome. For this reason
most work involving average Hamiltonians for RF pulse sequences
describes the spin Hamiltonian in terms of a spherical tensor basis, which is
analogous to the spherical tensor representation of rotations in physical
space. This representation will be only briefly introduced here—see the
more advanced texts in Further reading for more information.
In terms of irreducible spherical tensors, the homonuclear dipolar coupling
Hamiltonian of equation 4.25 is
The only non-zero missing image file hence the spin operator after
rotation by π is missing image file In other words, the homonuclear
dipolar Hamiltonian is unchanged by a 180° rotation. The relative ease of
expressing rotations in terms of irreducible spherical tensors is a major
incentive for their use when developing new RF irradiation schemes.
where missing image file and λ and λ ′ sum over the rank-1 and rank-2
terms respectively. Heteronuclear coupling terms are included in the rank-1
sum, since they contain only operators of the form Î z with respect to the
spins being manipulated by the RF.
where missing image file has been expressed in terms of the conventional
Cartesian spin operators, with missing image file This decomposition
effectively defines a new axis system. If the missing image file operator
in this frame is defined by
then the I -spin magnetization will precess about this z ′ axis. As discussed
further in section 5.5, z ′ does not usually coincide with the normal rotating
frame z axis, resulting in tilted axis precession . The scaling factor, α 1 ,
scales the rate of precession about this axis, and a small scaling factor will
result in poor spectral resolution. Hence the importance of maximizing α 1
while minimizing α 2 .
APPENDICES
Two of the appendices of the book develop themes that are relevant to this
chapter, but which are not required in later chapters. Appendix C introduces
the topic of Liouville space and how this is applied to problems involving
relaxation and site exchange. These are rather specialist topics and
understanding the theory behind them is not necessary in most applications
of solid-state NMR. Appendix D gives an introduction to the numerical
simulation of solid-state NMR spectra and experiments. Simulation is a
useful way of exploring the ideas developed in this chapter and is often
necessary in the development of new techniques and analysis of complex
problems in solid-state NMR. The software used for such simulations is,
however, subject to change and so the practical description of setting up
NMR simulations is confined to a distinct appendix.
FURTHER READING
“Spin dynamics ”, M.H. Levitt, John Wiley & Sons Ltd. (2008), ISBN 978
0 470 51117 6.
NOTES
2 The NMR shielding Hamiltonian, equation 4.11, also has this form, but
with the external magnetic field instead of the second spin.
3 D and equation 4.23 can be formulated in different ways, but these will
expand to the same overall expression for the dipole–dipole coupling
Hamiltonian.
10 The analytical derivation of the echo response is more involved than for
the simple spin echo and can be found in the Further reading. In contrast to
the simple spin echo, the solid echo cannot be pictured in simple vector
model terms. Moreover, the phase relationship of the two pulses is also
significant. While the relative phases of excitation and refocusing pulses
only affect the phase of the simple spin echo (the direction in xy along
which the magnetization is refocused), the phase difference must be 90° to
obtain a full solid echo.
14 Subtle issues such as active vs . passive rotations are being glossed over
here. See, for example, Levitt’s Spin Dynamics and Mueller’s Tensors and
Rotations in NMR (Further reading) for details.
15 Table B.2 of Schmidt-Rohr and Spiess in the Further reading contains
tables of Wigner functions.
17 The focus here is on the spin operators. Obviously the time dependence
of the spatial terms must also be included when determining average
Hamiltonians if MAS is involved.
CHAPTER 5
5.1 INTRODUCTION
The basic principles of solid-state NMR for dilute spin - nuclei, such as 13
C and 15 N, were set out in chapter 3 . This chapter first considers how the
basic techniques of Hartmann–Hahn cross- polarization (CP) and high-
power proton decoupling need to be modified under the conditions of high
magnetic fields and fast magic-angle spinning (MAS) rates which are now
available and which are increasingly important for the study of complex
systems, such as microcrystalline proteins.
Although straightforward one-dimensional NMR is usually the most
efficient route to solving chemical problems via solid-state NMR, there are
inevitably occasions where more “sophisticated” techniques, such as 2D
NMR, provide vital information that cannot be obtained from simple 1D
spectra. This chapter discusses more complex experiments involving
multidimensional NMR and/or homonuclear decoupling techniques (used to
improve 1 H resolution). The focus here is on experiments that provide
answers to essentially qualitative questions, such as “Is site A close to site
B?” The more demanding quantitative experiments, for example “What is
the distance between site A and B?”, are discussed separately in chapter 8 .
5.1.1 SPIN- NMR AT HIGH MAGNETIC FIELDS
Sideband m atching
5.1
Much of the power and flexibility of modern solution-state NMR flows from
the complementarity of information provided by chemical shifts and
through-bond (indirect) couplings; the chemical shift is key to resolving and
identifying sites with different chemical functionality, while indirect
couplings (often loosely termed J couplings) provide direct information on
the chemical connectivity between the different sites.
The situation is rather different in solids. The dipole–dipole interactions
(particularly those involving 1 H) are much larger, and techniques such as
MAS and high-power proton decoupling are necessary to suppress the
dipolar interactions in order to achieve useful chemical shift resolution .
Even after this, the linewidths observed in solids are often too broad to allow
other indirect couplings to be directly observed, with the exception of
spectra involving heavy atoms where J couplings are naturally large; for
example, coupling constants between two directly bonded 195 Pt nuclei may
be ~9 kHz.
Fortunately, it is often possible to exploit J couplings in the solid state
even when they are not resolved in the spectrum. As is illustrated in Figure
5.6 , a spin-echo experiment (see section 4.4.2) refocuses the evolution
under inhomogeneous factors and is often sufficient to reveal evolution
under J couplings, either homonuclear or heteronuclear.5 This evolution
appears as an oscillation superimposed on the decay due to relaxation and
other homogeneous line-broadening factors; that is, the signal for a given
spin coupled to one other will fit to:
Figure 5.6 31 P NMR spectra of a molybdenum pyrophosphate after spin-echo periods of different
durations . The peak intensities are modulated by the 2 J pp couplings (which are not refocused by the
spin echo), and the solid lines are fits to equation 5.2. These couplings, and hence the corresponding
P–O–P angle in the pyrophosphate (P2 O7 ) units, are very different for the inner and outer pairs of
peaks, providing direct structural information. The spectral linewidths must be dominated by
inhomogeneous effects, since couplings as small as 10 Hz are measured despite linewidths being over
an order of magnitude larger. (Figure adapted from results published in Lister et al ., Inorg. Chem . 49
(2010) 2290.)
5.2
under the effect of the couplings. Note that the apparent couplings
(about 60 Hz) are scaled by the homonuclear pulse sequence; this scaling is
discussed further in sections 4.7.3 and 5.5.1. As explained in inset 7.8,
adamantane is a somewhat special case since molecular motion strongly
reduces the magnitude of the dipolar couplings, which means they can be
suppressed relatively effectively. However, with efficient decoupling
strategies, couplings can be resolved in more typical organic solids,
allowing an important class of solution-state NMR experiments to be applied
to these samples. The resulting correlation experiments are discussed in
more detail below.
missing image file couplings, which are vital to many classic
solution-state NMR experiments such as COSY and TOCSY, are at least an
order of magnitude weaker than There is little immediate prospect of
such small couplings being resolved in the context of typical rigid organic
solids. However, 2 J couplings are often accessible: missing image file
couplings measured across hydrogen bonds have been used to probe
intermolecular interactions, and a number of researchers have investigated
the dependence of missing image file couplings on bond angle in X–
O–X units (as in Figure 5.6 ).
Dipolar couplings are in general much larger than corresponding indirect (J)
couplings, and so it ought to be relatively straightforward to exploit dipolar
coupling to transfer magnetization between the spins before the
magnetization decays away. Indeed, for 1 H spins in typical organic solids, it
is sufficient to use a short mixing period of the order of milliseconds; z
magnetization is efficiently exchanged between the spins via the dipolar
couplings (i.e., spin diffusion). The rate of spin diffusion may be reduced by
MAS, but it is still effective even at the highest achievable spinning rates.
Transfer of magnetization via the dipolar couplings becomes more
difficult when the coupling network is weaker, as these couplings are then
efficiently suppressed by the MAS required for good spectral resolution. For
example, the rate of direct 13 C, 13 C spin diffusion under MAS conditions is
generally too slow to be useful even in fully 13 C labeled compounds.8 In
addition, the rate of spin diffusion is likely to depend strongly on the sites
involved—the greater the frequency difference between the sites (relative to
their linewidths), the slower the rate of spin diffusion. This complicates the
interpretation of correlation peaks, since their intensities are not simply
related to spatial proximity.
In order to get around this problem, RF pulse sequences can be used to
recouple the dipolar interactions of interest, while using MAS to suppress the
unwanted interactions. In the absence of RF, evolution under individual
dipolar couplings is refocused over a rotor period, forming rotary echoes at
multiples of the rotor period (see Figure 3.15 and section 4.6.1). Unless the
coupling is sufficiently large for spinning sidebands to be observed, it is not
then possible to observe or exploit the coupling. However, addition of RF
pulses, synchronized with the rotation, disrupts the averaging by MAS. Fully
disrupting the MAS averaging to reintroduce CSAs as well as dipolar
couplings is obviously counterproductive, and so the RF pulse sequences are
designed to selectively recouple the interactions of interest.
Many such sequences have been devised, for recoupling both
homonuclear and heteronuclear interactions. For example, radiofrequency-driven
recoupling (RFDR) involves applying a train of rotation-synchronized π pulses
to recouple (see Figure 3.12 for an example). More complex schemes such
as POST-C7 are designed to create a well-defined “average Hamiltonian”
over the rotation period that depends only on the dipolar couplings. (Note
that such sequences often require fixed relationships between the spin rate,
and the relevant RF nutation rate, missing image file for example, the
condition missing image file must be met for C7-based sequences.) In
practice, the particular sequence used is not critical for the qualitative
applications described here.
For most applications, it is desirable for these sequences to be
“broadband,” that is, the efficiency of the dipolar recoupling should be
approximately constant across the spectrum. This ensures that the intensity
of any cross peak in a correlation experiment provides a reasonable measure
of the strength of the coupling between the pair of spins involved. The
magnetization transfer will be fastest between spins with the strongest
couplings, but as the mixing period increases, magnetization will equilibrate
between all the coupled spins. Such nonselective transfer is usually
appropriate for correlation experiments. Site-selective exchange of
magnetization is usually more relevant when the goal is measurement of
specific internuclear distances. These applications are discussed in section
8.3.
Figure 5.9 . (Left) Schematic of an H/X HETCOR experiment: 1 H magnetization evolves at the 1 H
NMR frequencies during t 1 , before being transferred (usually via CP to the X spins) where it is
detected. (Right) An 1 H/29 Si HETCOR spectrum of octosilicate acquired in about 14 h. In this case,
the 1 H spins are sufficiently weakly coupled to each other that 1 H homonuclear decoupling during t
1 was unnecessary. The traces at the top and l eft-hand side are separate 1-D spectra, not projections.
Figure 5.10 . 1 H/13 C HETCOR experiment on 3-methoxy benzoic acid using cross-polarization
contact times of (bottom) 100 μ s and (top) 400 μ s. At short mixing (contact) times, correlations are
only observed between protonated carbon sites and their associated H resonances. At longer times,
additional correlations are observed, for example, between the carboxylic acid resonance (at ~173
ppm, split by crystallographic inequivalence) and the acid protons (~13 ppm). Experiment time: ~14 h
for each spectrum.
Figure 5.14 . 31 P DQ/SQ homonuclear correlation experiments on a pyrophosphate material. (a) The
correlations in the J-based experiment (refocused INADEQUATE) show which pairs of resonances are
part of the same P2 O7 unit. The diagonal peak (*) is the result of a “rotational resonance” and is not
the result of J coupling (see text). (b) The dipolar-based experiment (using POST-C7 recoupling)
shows the correlations within the pyrophosphate units, and also weaker longer range couplings which
reflect interactions between nearby P2 O7 units. Dashed circles mark absent correlations and arrows
mark auto-correlation peaks (see text). (Figure adapted from results published in Lister et al ., Inorg.
Chem . 49 (2010) 2290.)
Interpretation of c ross p eaks in c orrelation s pectra
In most cases, the presence of a cross peak (or pair of cross peaks in double-
quantum/single experiments) can be interpreted straightforwardly in terms of
close spatial proximity of the spins involved (dipolar-based experiments) or
in terms of chemical bonding (experiments based on indirect (J) couplings).
There are, however, some subtle issues that must be borne in mind when
analyzing correlation spectra.
Figure 5.15 illustrates schematically the different types of DQ/SQ and
SQ/SQ correlation experiments. The information content of the double-
quantum and single-quantum variants is essentially the same (and the same
recoupling sequences can be used for both). The DQ/SQ spectra are, in
general, less crowded since peaks along the diagonal are not usually present
(but see the discussion of autocorrelation peaks below), whereas diagonal
peaks are always present in simple SQ/SQ correlation spectra.12 This makes
it easier to observe cross peaks between sites of similar NMR frequency.
In the solid state, two sites are only truly equivalent if they are related by a center
of inversion or simple translation; in this case, all the tensors for the NMR
interactions are always identical. If, however, there is a symmetry relationship
between the sites which is less than this, then only the isotropic components of the
NMR interactions are strictly identical. The sites will have the same isotropic
NMR frequency, but will be inequivalent at a general crystallite orientation as a
result, for example, of CSA. Coupling between such sites will result in a so-called
rotational resonance under sample spinning, leading to unusual NMR
lineshapes and unexpected autocorrelation peaks in 2D spectra (see Figure 8.7 for
an example of a related type of rotational resonance). Such rotational resonance
effects can be suppressed by “spinning out” the anisotropic interactions, but they
can, with care, be used to elucidate quite subtle questions of crystal symmetry.
5.5 HOMONUCLEAR DECOUPLING
The suppression of the strong dipolar couplings between protons in the solid
state has been a repeated theme in many of the experiments described above.
This section explores this important topic in greater detail.
As discussed in section 2.7 and inset 4.5, the Hamiltonian for a
collection of spins coupled by homonuclear dipolar couplings behaves
“homogeneously” under MAS; in contrast to other anisotropic interactions
such as the CSA, MAS does not break up the static lineshape into a set of
sharp spinning sidebands, but rather the sidebands (and centerband) have a
finite width that decreases only modestly with increasing spin rate. (To a
good first approximation, the linewidths decrease inversely with increasing
spin rate.)
The problem of limited resolution is particularly acute in 1 H NMR.
This is firstly due to the strength of the homonuclear interactions; 1 H has a
high magnetogyric ratio and individual dipolar couplings are large (e.g., over
20 kHz for the two protons in a rigid CH2 group), resulting in a large overall
homogeneous linewidth (typically ~50 kHz for a rigid solid). Secondly, the 1
H chemical shift range is limited to about 15 ppm, which corresponds to a
range of frequencies of only 7.5 kHz at an 1 H Larmor frequency of 500
MHz. As a result, even the fastest MAS is typically only able to resolve a
handful of distinct resonances. The situation is much easier for other
abundant spins such as 19 F and 31 P since overall coupling strengths are
lower and the range of NMR frequencies spanned is much larger,
particularly at higher magnetic fields. It is usually much easier when
working at current magnetic field strengths14 to use fast MAS rather than
homonuclear decoupling for these nuclei. The remainder of this section
focuses exclusively on homonuclear decoupling applied to 1 H.
Achieving useful resolution in a 1 H spectrum requires the effects of the
homonuclear couplings to be suppressed by more than two orders of
magnitude. This is a demanding requirement, and even small experimental
imperfections can become noticeable when such high levels of performance
are needed. Moreover, homonuclear decoupling is subject to a variety of
experimental factors that are often hard to control or characterize (e.g.,
variation of B 1 across the sample, finite rise and fall time of pulses). The
higher the resolution expected from the homonuclear decoupling, the more
critical the experimental set-up.
A significant complication associated with all the widely used
decoupling sequences (see below) is that the 1 H magnetization precesses
about a tilted axis during the homonuclear decoupling, rather than precessing
about the z axis as is the case in the absence of RF irradiation (Figure 5.16 ).
Moreover, the exact position of the precession axis is not always well
defined, as a result of experimental imperfections. Sampling the raw x and y
magnetization leads to quadrature errors when the signal is Fourier transformed,
that is, artifact peaks appear at frequency –f for each genuine signal at
frequency f . The usual solution is to perform the experiment off-resonance;
the transmitter frequency is placed to one end of the 1 H spectrum making it
unnecessary to acquire a full quadrature signal. This may degrade the
effectiveness of decoupling for peaks furthest from the transmitter, but the
limited frequen cy range of the 1 H spectrum means that off-resonance
detection is usually preferred.
Figure 5.17 . DQ/SQ 1 H correlation experiments on a small dipeptide using (left) 30 kHz magic-
angle spinning (MAS) and (right) 12.5 kHz MAS + homonuclear decoupling in both dimensions. The
improvements in resolution are substantial (about a factor of 5). (Figure adapted from results
published in Brown et al. , J. Am. Chem. Soc. 126 (2004) 13230.)
with respect to the z axis. Magnetization precesses16 about this tilted axis
and the homonuclear couplings are suppressed (to first order).
Simple LG decoupling is not particularly competitive for achieving
high-resolution spectra since it only suppresses the dipolar line broadening
to lowest order. In order to suppress “higher order” terms and hence achieve
more effective decoupling, frequency-switched Lee-Goldburg (FSLG) decoupling
(Figure 5.18 ) alternates the sign of the frequency offset and switches the RF
phase. This significantly improves decoupling performance. Note that FSLG
is frequently implemented using continuous linear ramps of the RF phase,
avoiding the need for frequency switching, in which case it is usually termed
phase-modulated Lee-Goldburg (PMLG) decoupling.
Although there are close analogies to physical MAS of the sample, it is
important to note the differences. In MAS, it is the spatial component of the
Hamiltonian (the interaction tensor) that is being modulated by the spinning.
Hence isotropic components (such as the isotropic chemical shift) are
unaffected by the averaging process. In LG decoupling, however, it is, in
effect, the spin components of the Hamiltonian that are being modulated.
“Bilinear” interactions such as the homonuclear dipolar couplings are
averaged (to first order). But “linear” terms, such as the chemical shift
Hamiltonian or heteronuclear couplings, are also partially averaged—they
are scaled by a factor of missing image file for LG decoupling. Hence
if a spin would precess at frequency Ω in the absence of RF, it will precess at
a frequency missing image file under LG decoupling.
As discussed in more detail in section 4.7.3, this scaling of the
precession frequencies is an unavoidable consequence of using RF
irradiation to suppress a component of the spin Hamiltonian, and it is
important to correct for this scaling before extracting chemical shift
information from CRAMPS spectra. Although the ideal scaling factor for a
given sequence will be known, observed scaling factors are quite sensitive to
experimental factors, and it is good practice to measure the actual scaling
factor on a known set-up sample.
FURTHER READING
2D AND MULTI-DIMENSIONAL NMR APPLIED TO THE SOLUTION
STATE
“Understanding NMR spectroscopy” , J. Keeler, Wiley (2005), ISBN 978 0 470 01786 9.
“High-resolution NMR techniques in organic chemistry” , T.D.W. Claridge, Elsevier (2009), ISBN
978 0 080 54818 0.
HETERONUCLEAR DECOUPLING
“Heteronuclear decoupling in the NMR of solids”, P. Hodgkinson, Prog. NMR Spectry. , 46 (2005)
197. DOI: 10.1016/j.pnmrs.2005.04.002.
BIOMOLECULAR APPLICATIONS
“NMR spectroscopy of biological solids”, Ed. A. Ramamoorthy, CRC Press (2005), ISBN 978 1 574
44496 4.
NOTES
1 Spinning sideband suppression techniques, as discussed in section 3.3.6, are only a partial solution
as they cancel out the sidebands without increasing centerband intensity.
2 Faster MAS also decreases 1 H spin diffusion rates (see section 2.7), slowing the rate at which
magnetization can be transferred from more distant 1 H spins to the X spin. This further diminishes
overall CP efficiency.
3 At first sight, the time dependence of the matching condition in ramped CP might be expected to
reduce the effectiveness of polarization transfer at an individual matching condition. In fact, sweeping
through a match can (in principle) result in a more complete (adiabatic ) transfer of polarization from
one spin to the other.
4 It is a common misconception that T relaxation must be fast in solids. Classic NMR relaxation
2
theory assumes that the motional processes occur on a much faster timescale than the NMR frequency
and so its results cannot be simply extrapolated to the slow-motion limit; indeed T 2 relaxation must
be slow in the absence of motional processes to drive it. At the other end of the scale, the only
difference between a plastic crystalline phase and a liquid is the absence of translational motion.
Hence it would be expected that missing image file for such “soft solids” as for the fast-motion
limit in solution-state NMR.
5 The heteronuclear case requires simultaneous π pulses on both nuclei.
6 Generally, the amplitude of the signal is modulated at the start of t , rather than its phase . “Phase
2
modulation” usually results in “twisted” lineshapes in 2D spectra, which cannot be phased into all-
positive peaks. These subtle, but important, issues in 2D NMR are discussed in more detail in the
Further reading.
7 In some cases, it can be useful to “open up” the filters and deliberately allow extensive folding in the
direct dimension, for example to superimpose all the peaks of a spinning sideband manifold.
8 Carbon-carbon proximity information must be obtained via effi cient 1 H spin diffusion. Such
experiments, such as proton-driven spin diffusion , are only viable for labeled compounds and are not
considered further.
9 The choice of probe and MAS rotor for heteronuclear experiments is generally determined by the
least sensitive nucleus. As this will be the dilute spin (for unlabeled samples), the rotor diameter will
tend to be relatively large, limiting the maximum MAS rate that can be used.
10 “Refocused” refers to the refocusing period after t which converts the anti-phase doublets
1
normally observed in solution-state INADEQUATE experiments into in-phase doublets. The lower
resolution in typical solid-state NMR spectra makes the acquisition of anti-phase doublets highly
undesirable, since the overlap of adjacent signals with opposite signs results in extensive signal
cancellation.
11 The original solution-state 13 C INADEQUATE experiment is rarely used due to its extremely low
sensitivity (the signal arises from the tiny fraction of C–C bonds in which both nuclei are 13 C).
Although an even more demanding experiment in the solid state, it may be the only means to
completely assign a 13 C spectrum. The situation is considerably eased if samples can be globally
enriched in 13 C.
12 With appropriate phase cycling, it is possible to modify SQ/SQ correlation experiments so that only
pathways involving double-quantum coherences are selected. The information content of such double-
quantum-filtered experiments is then equivalent to the DQ/SQ experiment, with the only significant
difference lying in the form of the spectra.
13 The reason for the difference lies in the different form of the Hamiltonians for the rank-0
(isotropic) and rank-2 tensor components of NMR couplings (see section 4.2.3). Figure 2.4 shows the
spectrum of a pair of equivalent spins coupled by dipole–dipole coupling (pure rank-2); the equivalent
spectrum for a pair of identical spins coupled by an isotropic J (indirect dipole–dipole) interaction
(pure rank-0) would be a simple singlet!
14 However, increased magnetic fields also mean that the contribution of CSAs and isotropic chemical
shift differences to the overall Hamiltonian can become substantial for nuclei such as 19 F. As a result,
decoupling sequences derived for 1 H NMR often perform poorly when applied to other nuclei.
15 Windowless sequences (which use continuous RF) will tend to decouple more efficiently than
windowed sequences consisting of isolated pulses. Indeed, recent work that applies homonuclear
decoupling in directly acquired dimensions has generally inserted acquisition windows into new
windowless sequences (such as PMLG, DUMBO) rather than using classic windowed sequences.
16 Precession rather than nutation is used here as the spin evolution over complete cycles of the Lee-
Goldburg irradiation rather than the rapid nutation during the LG cycle is being considered. See the
discussion of WHH-4 decoupling in section 4.7.1 for more details.
17 Note that the effective nutation rate under off-resonance irradiation is given by
missing image file which corresponds to an increase by a factor of missing image file at the LG
condition. The matching condition needs to be adjusted accordingly.
CHAPTER 6
QUADRUPOLAR NUCLEI
6.1 INTRODUCTION
Figure 6.1 . Simulated magic-angle spinning (MAS) spectra for a first-order spin - nucleus with χ = 150 kHz:
(a) η = 0, and (b) η = 1. Shielding anisotropy is ignored. The intensity of the central transition is truncated.
Figure 6.2 . Deuterium magic-angle spinning (MAS) spectrum at 7.05 T and a spin rate of 8 kHz for 2,3-d2 -
fuma ric acid. The quadrupolar parameters are χ ≅ 160 kHz and η ≅ 0.
The strong central feature of spectra for nuclides with I an odd multiple of can
be readily used to determine the isotropic chemical shift for both static and MAS
cases (but see later for complications in cases with large quadrupole coupling
constants). For nuclides with integral spin, such a measurement is more difficult and
it may be necessary to take the full bandshape into account. However, in both cases,
it is feasible to fit the total bandshape to give the quadrupole coupling constant;
suitable computer programs are available for simulating and iteratively fitting both
static and MAS spectra, in the latter case using the intensities in the spinning
sideband manifolds. The results, in turn, give information on the electronic
environment of the nucleus and thus on the bonding situation. Quantum theory is
often used to compute quadrupole coupling constants from crystal structure
information, thus providing a sensitive test for both theory and experiment.
Thus far, it has been tacitly assumed that axially symmetric sites are being
discussed. However, in general, the quadrupole coupling tensor is asymmetric.
Asymmetries affect the appearance of static spectra (and hence of MAS spectra
also). Figure 6.1 (b) shows an example for a spin - nucleus with η = 1 , for
comparison with the η = 0 case of figure 6.1 (a). The computer programs mentioned
above take account of asymmetries and thus these parameters can be derived by
fitting static spectra or spinning sideband manifolds.
The accurate setting of the magic angle is particularly important for
quadrupolar nuclei because of the frequently very wide range of the resonance. A
very small offset from the magic angle is immediately manifested in line broadening
for the spinning sidebands of the satellite transitions. In fact, the usual way of
setting the magic angle is to optimize the linewidths of spinning sidebands for the 79
Br spectrum of KBr. The optimization is best done by observing the duration of the
free-induction decay (FID), as shown in figure 6.3 .
Figure 6.3 . Bromine-79 free-induction decays (FIDs) of KBr, spinning at 3 kHz (a) off (by <1° ) and (b) on the
magic angle at a magnetic field of 9.4 T. The insets show portions of the corresponding spectra away from the
centerband.
The total intensity in a given subspectrum (central or satellite transition) is
governed by the raising and lowering operators (I + and I – , respectively; see section
4.2.1) and is, therefore, for the transition proportional to:
6.1
I Ratio
3:4:3
5:8:9:8:5
7 : 12 : 15 : 16 : 15 : 12 : 7
9 : 16 : 21 : 24 : 25 : 24 : 21 : 16 : 9
Figure 6.4 . Energy-level diagram for a spin- nucleus in a magnetic field B 0 , showing the effects of zero
(left-hand side), first-order (center), and second-order (right-hand side) quadrupolar perturbations, the last
leading (in general) to a decrease in the resonance frequency for the central transition.
6.4
where and are the raising and lowering operators, respectively (see section
4.2.1).
Before this equation is developed here, the reasons why the results are often
mathematically complicated should be understood. The first is that all Hamiltonian
terms must be given in the same axis system. However, equation 6.4 is expressed in
the principal axis frame of the EFG tensor, whereas the Zeeman energy is
determined by B 0 , which lies in the laboratory frame of reference. Therefore, a
conversion of the spin operators from one frame to the other is necessary, which
involves rotations, generally over two angles (see section 4.6 for details). In fact,
when MAS is employed, the rotor axis must also be considered and an additional
rotation of axes is necessary. Converting Hamiltonian terms by rotations is relatively
simple when the tensors are axial, since then only one angle is involved (e.g., that
between the Z EFG axis and B ).5 Thus, when asymmetry is neglected, equation
0
6.4 becomes, in the laboratory frame:
6.5
This is the origin of the familiar 3cos2 θ – 1 factor, which is averaged to zero by
MAS (though proper consideration involves transformation to the axes of the
rotation). However, the angular dependence in the second and third terms within the
square brackets cannot be averaged by MAS and is the source of difficulties for
high-resolution spectra of quadrupolar nuclei, as discussed in the next section.
Significant additional complications result if the quadrupole tensor is asymmetric,
since three angles are then needed to describe rotations of axes.
The second reason for complications lies in the nature of the spin operators.
Terms such as are secular (i.e., affect the energies of states but do not mix
states significantly); application of such operators gives first-order energy changes.
Terms such as missing image file and missing image file (se e sec tion
4.2.1), on the other hand, mix states (in the sense that the wavefunctions are
significantly changed), requiring substantial mathematical manipulation. This gives
rise to second-order contributions to the energies.
In what follows in this section and later sections, the text will deal with various
relatively simple situations, starting with the effects of asymmetry on first-order
spectra.
When asymmetry is included, the value of missing image file in the
laboratory frame is found to be:
(Capital letters for the axes indicate the principal components, as usual.) If only the
secular terms are retained (i.e., at relatively low quadrupole coupling), the truncated
quadrupolar Hamiltonian becomes:
Figure 6.5 shows schematically the effect of asymmetry on the static powder pattern
for a spin - n ucleus. The two subspectra for the transitions missing image file
and missing image file are indicated. In all cases the spectrum is symmetrical (in
the absence of shielding anisotropy). Thus the subspectra for the satellite transitions
are mirror images for all values of the asymmetry parameter; they completely
overlap when η = 1.
energy decreases and spectra become less complicated by the second-order effects
to be described below. Generally speaking, this results in better resolution
(enhanced because chemical shifts increase in frequency units). There are also gains
in signal intensity with increase in B 0 .
When quadrupole coupling constants are significant with respect to the Zeeman
term, it becomes necessary to include perturbation to second order, which causes a
number of complications to appear in the spectra. This is because terms involving
the raising and lowering operators cause mixing of states, as mentioned above,
which means that the eigenstates are no longer pure Zeeman states. This leads to
second-order energy contributions and thence to shifts in the spectra. The evaluation
of the relevant equations can be tedious unless advanced methods (spherical
harmonics) are used (see chapter 4 ), and the results are complicated. Here only an
equation for the second-order quadrupolar energy, missing image file of a static
sample in an axially symmetrical case will be given:
where θ is the angle between the EFG symmetry axis and B 0 , while ν 0 is the
Larmor frequency, and m I is the spin component quantum number of the energy
level in question. In equation 6.9, the three terms in the square brackets are referred
to as zeroth-, second-, and fourth-rank contributions . They involve the zeroth-,
second-, and fourth-degree Legendre polynomials , respectively. The zeroth-de gree
polynomial is unity and the other two are:
These two factors arise from the requirement to rotate the spin operators between
different axis systems (in this case those of B 0 and the EFG). The second-degree
polynomial will be recognized as the one involved in MAS , as discussed in chapter
2.
Since equation 6.9 depends on m I as well as missing image file the
energies of the levels missing image file are changed differentially. In fact the
proportionality of E Q to mI shows that they are affected to equal but opposite
extents. Therefore, the splitting between these levels (the central transition) is
altered (in contrast to the zero effect of the first-order quadrupolar term)—see figure
6.4 . Moreover, the magnitude of this effect depends on the orientation of the EFG
as expressed by θ . Therefore, the central transition will be a powder pattern when
polycrystalline or amorphous materials are studied.
The simplest form for the resonance frequency of the central transition (
missing image file ) for an axially symmetric case, which can be derived from
equation 6.9 , is:
There is a general shift away from ν 0 for all orientations except for θ = 0 ° ,
180 ° , and missing image file ( θ = 70.53 ° and 109.47 ° ).
Since this shift depends on the orientation of the EFG in B 0 , a
microcrystalline sample will give a powder pattern, in contrast to the first-
order perturbation (which left the central transition as a single line).
The center of gravity of the powder pattern will not be at ν 0 .
The effect is inversely proportional to ν 0 and will therefore become smaller if
B 0 is increased.
The magnitude of the effect decreases with I , other things being equal (see
inset 6.2).
Since the second-order term as a whole is not proportional to
missing image file , MAS cannot eliminate the effect.
Inset 6.2. Dependence of the second-order effect on the nuclear spin quantum number
for the central transition
The value of the spin-dependent factor, f ( I ), is as follows:
This factor means that, for comparable quadrupolar coupling constants, second-order
effects get significantly smaller as I increases, which is perhaps counterintuitive. Thus,
second-order broadening of the central transition for 87 Sr ( missing image file ) is only
0.055 that of 87 Rb ( missing image file ) in spite of the fact that the former has the larger
quadrupole moment (33.5 vs. 13.35 fm 2 ).
Figure 6.6 . Central-transition 27 Al spectra of polycrystalline aluminum acetylacetonate. Top: static sample.
Bottom: Sample rotated at the magic angle at 9.26 kHz. The position of the isotropic chemical shift is indicated
as ν 0 . The small peak just to high frequency of ν 0 arises from the centerband of the inner satellite transition (
missing image file ).
While equation 6.12 is useful for descriptive purposes, greater insight can be
obtained by recasting it into a form obtained by use of Legendre Polynomials (see
equation 6.9):
The center of gravity of the powder pattern for the central transition is determined
by an isotropic second-order shift , given (for zero asymmetry) by the first term of equation
6.14:
Thus, as mentioned earlier, the center of gravity is shifted to low frequency, though
(for a static sample) the powder pattern extends both sides of the isotropic shielding
position.
6.4.3 CENTRAL-TRANSITION SPECTRA: RAPID SAMPLE SPINNING
When a sample is spun about an angle β to B 0 , there is some (though not complete)
averaging of the second-order contribution to the resonance frequency. Equation
6.14 shows that the average values of missing image file and missing image file
and are required. This involves another change of axes and therefore brings in new
Legendre polynomials. The necessary average of equation 6.14 can be shown to be:
where the angle β is that between the rotor axis and B 0 , while missing image file
is the angle between the Z axis of the EFG and the rotor axis (which can take any
value for a powder sample). The factors missing image file and
missing image file act to scale the relevant terms for the powder bandshape. At the
magic angle, missing image file the middle term is eliminated, while the final term
is scaled by missing image file
Thus the second-order shift under conditions of MAS (including the isotropic
part) becomes:
This term describes a powder pattern, with low-frequency shifts for all values of
missing image file (unlike the static case). The center of gravity ( equation 6.15 )
is invariant to the spinning. The lower spectrum in figure 6. 6 shows the case of
aluminum acetylacetonate under MAS conditions. Clearly, the second-order effects
depend on the ratio missing image file and thus inversely on the magnitude of the
applied magnetic field B 0 . Figure 6.7 shows such a variation of the central
transition for a zero-asymmetry case. The bottom spectrum contains sharp spinning
sidebands from the outer transitions, which become broadened as
missing image file increases. At higher missing image file ratios, spinning
sidebands of the central transition appear.
Direct measurements from spectra such as that shown in figure 6.6 can yield
information about both the isotropic shielding and the quadrupolar coupling
constant, the latter being strongly related to the electronic environment of the
nucleus in question. Inset 6.3 indicates how the relevant parameters may be derived
from the MAS powder pattern of the central transition for an axially symmetric
case. Optimum choice of operating magnetic field depends on the aim of the
experiment. Resolution is best for the highest B 0 (and this makes determination of
isotropic chemical shifts relatively easy). However, when quadrupole coupling
constants are small, it may be advantageous to obtain the quadrupolar parameters
from the central-transition bandshape at relatively low applied magnetic field so as
to increase the magnitude of the second-order broadening. Alternatively, such
parameters may be derived from analysis of the spinning sidebands from the
satellite transitions at low missing image file .
Figure 6.7 . Variation of the central-transition region of a spin - spectrum (simulated) with the ratio of the
quadrupolar coupling constant to the Larmor frequency , missing image file . The dotted line indicates the
isotropic chemical shift.
Inset 6.3. Analysis of the central transition for a second-order MAS powder pattern
Figure 6.8 shows a schematic powder pattern for the central transition of a quadrupolar
nucleus under MAS conditions, with the positions of the three turning points indicated.
Expressions for these turning points may be obtained by differentiating equation 6.17.
They are found to be at missing image file = 0 ° , 90 ° , and 49.1 ° (with values of 5 –
18cos 2 missing image file + 21 cos 4 missing image file equal to 8, 5, and
missing image file , respectively). The positions are given in figure 6.8 in units of K ,
where:
For a zero-asymmetry case with a well-defined powder pattern, the quadrupole coupling
constant can be readily derived from the splitting between the prominent horns of the
powder pattern, which is missing image file In the case of aluminum acetylacetonate (see
figure 6.6 ), this splitting is 670 Hz, leading to a value of 2.7 MHz for missing image file
. The powder pattern may be reproduced in full by suitable computer programs and the
values of missing image file and missing image file obtained more accurately by
iterative fitting of the full bandshape. For the aluminum acetylacetonate case, a value of 3.0
MHz for missing image file was obtained from a fit of the full bandshape. Although
assumed to be zero in the approximation above, the fit gave an asymmetry value of
missing image file
missing image file
Figure 6.8 . Schematic magic-angle spinning (MA S) NMR powder pattern for the central transition of a second-order
spectrum for a quadrupolar nucleus with an axially symmetric EFG. The angle missing image file = 22. 2 ° gives rise
to a resonance at the same frequency as the turning point at 90° .
The relevant equations for cases with missing image file are significantly more
complex and will not be quoted here. The reader is referred to other texts, as listed
in the Further reading section, for details. However, the general effects of
missing image file for the MAS case are shown in figure 6.9 . For a few purposes,
the effect of asymmetry is simply introduced by a factor of missing image file .
For example, in equation 6.15 for the isotropic second-order effect, the factor 3 must
be replaced by missing image file . Since missing image file , the influence of
this factor in such cases is rather small.
Of course, real examples do not often look like the ideal cases of figure 6.9 .
Thus figure 6.10 shows the central-transition portion of the 27 Al ( I =
missing image file ) spectrum for a natural aluminosilicate clay. The
relatively featureless spectrum is typical of those from many aluminosilicate
materials that contain some disorder. The two intense centerband resonances are
from aluminum octahedrally ( δ Al ≅ 1 ppm) and tetrahedrally ( δ Al ≅ 58 ppm)
coordinated by oxygen. The remaining signals are spinning sidebands. While the
theoretical shapes of the resonances are not apparent, the isotropic chemical shifts
(including the second-order quadrupolar contribution) can be derived from the
centers of gravity of the signals. If the spectrum is measured as a function of
magnetic field, the true isotropic shift (i.e., excluding quadrupolar effects) can be
obtained by plotting the center of gravity as a function of the inverse of the Larmor
frequency and extrapolating to zero.
For most samples, spin–lattice relaxation times of quadrupolar nuclei are short since
they are dominated by the quadrupolar mechanism and this is normally efficient (a
consequence of the large quadrupolar interactions!) provided there is some
molecular-level mobility. Therefore direct-excitation methods (repeated single
pulses, each followed by an FID) are normally used (for both central and satellite
transitions). Recycle delays following the FIDs do not need to be lengthy (in
contrast to the case for spin- nuclei). This implies that the optimum signal-to-noise
ratio may be obtained from pulse angles less than 90° , so it may be worth using the
Ernst angle (see section 3.3.1). Moreover, there are other reasons for keeping the
pulse angle low (see below). Proton decoupling may or may not be required,
depending on the chemical system involved.
An example of the potential value of obtaining quadrupole coupling
information is provided by the 23 Na spectrum of a mixture of sodium chloride and
sodium nitrite shown in figure 6.11 . It is immediately obvious that the high-
frequency signal has a negligible quadrupole coupling constant and so the sodium
atoms must be at a site of very high symmetry. This establishes the assignment as
that of sodium chloride. The other signal, which must be from sodium nitrite, has
the typical appearance arising from an axially symmetric site; simulation confirms
this and gives missing image file = 1.09 MHz. This is consistent with the
crystal structure and would have given structural information in the absence of
diffraction data.
6.6.1 NUTATION
However, the existence of large quadrupole coupling constants raises new issues for
obtaining spectra . During solid-state experiments on spin- nuclei, the RF pulse
energy E RF (generally expressed as the equivalent angular frequency, γ B 1 ) is
sufficiently high in relation to the strengths of dipolar and shielding interactions that
they can be ignored during the pulse. However, that is not necessarily the case for
the quadrupolar interaction. In the situation where the quadrupolar energy is much
less than the RF energy, missing image file , nutation (rotation of the magnetization
under the influence of the RF; see section 3.2 .1) will occur as in a case with zero
quadrupole coupling, that is at an angular frequency γ B 1 . This is known as a hard
pulse and results in all the resonances (central and satellite transitions) being
observed.
On the other hand, in the extreme case that missing image file nutation
for the central-transition magnetization occurs at an angular frequency
missing image file where missing image file is known as the Rabi factor.
In such a case, the RF duration required for a 90° pulse is shorter than that for an
equivalent situation for a solution; for an missing image file case, for
example, the Rabi factor is 2. Such a situation is said to involve a soft pulse and the
effect is selective, which means that only the central transition of the quadrupolar
energy-level system will be fully affected. Both soft pulses and hard pulses have
uses for quadrupolar spectra. For intermediate values of the pulse power (sometimes
also referred to, rather confusingly, as “hard,” but better called simply nonselective ),6
the nutation situation is complex, and different coherences (see section 4.3) are
created. Figure 6.13 shows the development of two such coherences, namely single
quantum (central transition, SQ) and triple quantum (TQ) for a spin-
missing image file system following such a pulse. Because TQ coherence
cannot be created directly but only via the creation of SQ coherence, there is an
apparent “induction period” prior to the appearance of the former, as can be clearly
seen in figure 6.13 .
Of course, a particular pulse may be hard in respect of some nuclei but soft in
relation to other nuclei of the same isotope in the same sample. Variation of the
pulse duration, with two-dimensional Fourier transformation, gives a nutation spectrum
that can sharply reveal differences of quadrupolar interaction strengths in cases
where resonances overlap in one-dimensional spectra. However, such differences
produce problems for determining relative intensities between signals in well-
resolved one-dimensional spectra. The easiest way to obtain good results
semiquantitatively is to employ small pulse angles (≤8 ° ), since deviations between
the various cases are relatively small under such conditions. However, this does not
optimize the signal intensity for any of the nuclear sites.
Although it can be useful for the derivation of quadrupolar parameters, the broad
bandshapes caused by the second-order quadrupolar effects introduce problems of
spectral resolution, making analysis complicated if there is more than one relevant
site for the nucleus in the crystals in question. In principle, it would be useful to be
able to eliminate the second-order effects so as to improve resolution and to obtain
chemical shifts more readily. One obvious method is to use high magnetic fields, but
these are frequently not accessible and, in many cases, spectrometers operating at
sufficiently high fields do not exist. There are several more sophisticated ways
round this problem by manipulating the quadrupolar effect, as described in the next
few subsections. However, these techniques are largely confined to cases with
relatively small quadrupole coupling constants since impossibly high spin rates
would be required otherwise.
This technique provides two-dimensional spectra (see section 5.4.1 for a general
discussion of 2D operation), giving isotropic spectra in one dimension and
quadrupolar-broadened anisotropic spectra in the other. The experiment consists of
spinning at different angles during two times t 1 and kt 1 , where the constant k is
calculated so as to provide the elimination of the anisotropic effects. A CT -selective
90° pulse is applied prior to each of these times and the magnetization is stored in
the z direction (by another 90° pulse) between them while the rotor angle is changed.
The FID is recorded during time t 2 following the end of kt 1 . As is usual in two-
dimensional experiments, t 1 is incremented and the dataset S (t 1 , t 2 ) is doubly
Fourier transformed. In the f 1 dimension, the second-order quadrupolar broadening
is eliminated (though the second-order isotropic effect remains). There are many
pairs of angles that can be used, but none eliminate shielding anisotropy or dipolar
interactions, so resolution remains a problem. The best pair of angles is 0° and 63°
(which requires k = 5) , since the latter is not far from the magic angle (so that in t 2
the first-order anisotropic effects are significantly reduced). This method is also
technically demanding, since it requires rapid changes of rotor angle, but (in
contrast to DOR) there is no difficulty about obtaining high spin rates because only
one rotor is involved.
The effect of the first pulse depends on the size of the quadrupole coupling, so
it is sample and environment dependent. The implications of this, for samples
containing more than one environment, will be discussed later. After an evolution
time, t 1 (which can be rotor synchronized although this is not essential), the ±3
coherence order is converted to an observable –1 coherence order with the second
pulse. As usual for two-dimensional experiments, FIDs are collected over a time t 2
for various values of time t 1 to produce a two-dimensional data set involving two
time variables. Double Fourier transformation then gives a two-dimensional plot
involving two frequency dimensions, f 2 and f 1 , respectively. The direct dimension
(frequency f 2 ) may be labeled MAS or 1Q since a projection onto this axis results
in the normal one-dimensional spectrum. The indirect dimension (frequency f 1 ) is
labeled MQ (or, specifically, 3Q, 5Q, etc., as appropriate). The two coherences ( p =
±3) selected by the first ( excitation ) pulse are produced with equal efficiencies.
However, the second ( reconversion ) pulse, which is arranged to give observable p = –
1 coherence, acts unsymmetrically, that is the effectiveness of the
missing image file and missing image file coherence transfer steps is
different. If the duration of this pulse is chosen badly, then there will be an
imbalance between the signals generated by each pathway, which results in a poor
final lineshape. Conversely, for the missing image file case only, if the
duration of the pulse is chosen so that the two transfer steps produce an equal
amount of signal then a pure absorption lineshape will result. For higher spins, a
pulse duration can be selected to give a good approximation to a pure-phase
spectrum, as shown in figure 6.17 (b)9 for a spin - missing image file case.
Figure 6.17 . Triple-quantum 27 Al MQMAS spectra from aluminum acetylacetonate. The spectra were
obtained at 9.4 T, with an excitation-pulse duration of 3.6 μ s at an RF field equivalent to 110 kH z. The
reconversion pulse durations were 0.8 and 2.0 μ s for (a) and (b), respectively. (a) Contour plot illustrating the
poor lineshape (containing a large dispersive component) obtained when the two coherence transfer pathways
are combined unequally. The dashed lines represent negative intensity. (b) The same experiment but with
coherence transfer pathways combined approximately equally.
The resonance shown in figure 6.17 (b) does not lie parallel to the axis in the
directly detected dimension. This occurs because the second-order effects influence
the spectra in both dimensions. A process known as shearing is used to compensate
for this so that the resonance will then lie parallel to the MAS dimension (figure
6.18 ). This matter was not mentioned when multidimensional spectra were first
introduced (in chapter 5 ) because it is rarely important for spin-
missing image file spectra. The Fourier transform routines of the
spectrometer cope with shearing automatically. (This matter is described in detail in
the review articles listed in the Further reading section.) After shearing, one of the
axes is of mixed 1Q and MQ character. It has become common to label this as the
“isotropic” axis, though this is potentially confusing (see later in this subsection).
Figure 6.18 . The triple-quantum 27 Al MQMAS spectrum from aluminum acetylacetonate , as shown in Figure
6.17 (b) but after shearing has been applied.
Up to this point, much emphasis has been placed on the central transition. However,
as stated earlier, the complete static spectrum or the intensities of the satellite
transition spinning sideband manifolds (in particular) may be computer-fitted to
determine the quadrupolar parameters. Inset 6.6 gives the basic theory for the
satellite transitions. It shows that it is possible to identify the true chemical shift by
locating the various centers of gravity. This can be a simple process for spin-
missing image file if the inner and outer satellite peaks can be resolved. In this
case, the isotropic quadrupolar shifts for the outer, inner, and central transitions are
in the ratio 28 : 1 : –8 (figure 6.21 ), with a constant factor:
In order to achieve spectra of the type shown in figure 6.21 , the magic angle must
be accurately set and rotor synchronization must be used—or else all the spinning
sidebands must be carefully added into the centerbands (as was done in the case of
figure 6.21 ).
The center of gravity of the inner satellites is actually very close to the true
isotropic chemical shift. Moreover, the second-order broadening of transitions is
proportional to the relevant A 4 (I , mI ) coefficient listed in table 6.2 . For spin -
missing image file , these show that the widths for individual spinning
sideband signals for the outer, inner, and central transitions are in the ratio:
(The negative sign implies a reversal of the shape from that shown in figure 6.9 .)
The inner satellite lines are appreciably sharper than the others.
Equation 6.14 may be expanded to the general missing image file case:
where the spin-dependent coefficients are given (for spin- missing image file and spin-
missing image file nuclei) in table 6.2 . The above equation assumes axial symmetry.
Table 6.2. Spin coeffi cients for single-quantum transitions (see equation 6.19)
Table Missing
The analogous equation to 6.19 for sample spinning can be obtained by replacing P 2 (cos θ )
and P 4 (cos θ ) by P 2 (cos β ) P 2 (cos Θ ) and P 4 (cos β ) P 4 (cos Θ ) , respectively (as in going
from equation 6.14 to equation 6.16). Thus, under MAS conditions, equation 6.19
becomes, for the missing image file transition:
For the specific case of the missing image file satellite transitions, the quadrupolar shift
is:
Since the A 0 (I , mI ) coefficients differ for the central and satellite transitions, so will the
centers of gravity.
Figure 6.23 . Satellite transition magic-angle spinning (STMAS) 87 Rb spectrum of rubidium nitrate after
shearing. Operating parameters: the first two pulses were of duration 2.3 μ s and angle 90° (as measured for an
aqueous solution). The RF power was equivalent to 108 kHz. The z -filter pulse was of duration 6.0 μ s at a 14
kHz power equivalent. The spin rate was 10 kHz and the spectral width was 10 kHz in both dimensions.
While STMAS is somewhat more sensitive than MQMAS, it is significantly
more demanding to set up. The magic angle must be set precisely and the spin rate
must be highly stable.
Mention was made in section 2.6 that there are only four nuclides of significance for
NMR with integral spin quantum numbers. Of these, 10 B has I = 3, while the others
(2 H, 6 Li, and 14 N) are spin-1. Boron-10 NMR is unimportant because 11 B is the
preferred boron nuclide. Lithium-6 has a low quadrupole coupling constant and can
usually be treated as a “pseudo spin- missing image file ” nuclide—it provides a
useful foil to the spin- missing image file nuclide 7 Li. Consequently, only
deuterium and nitrogen-14 need to be considered in any detail here.
Quadrupole coupling constants for 2 H are generally modest (< ~300 kHz), so
second-order effects are not a problem and methods analogous to those used for
spin- nuclei generally suffice (though the spin dynamics is different). An example
has been given as figure 6.2 . Static bandshapes and spinning sideband manifolds in
2 H spectra frequently reflect local molecular-level mobility and are much used in
that context. Figure 6.24 shows an example (a 2 H MAS spectrum). The spinning-
sideband manifolds for finasteride hydrate THF solvate cover only ~35 kHz,
whereas the expected quadrupole coupling constants for a static molecule are ~200
kHz. There must be rapid (though not isotropic) motion of the THF molecule in the
solvate. However, the quadrupole parameters for the two sites are still unequal—for
the high-frequency peak χ = 20.2 kHz and η = 0.99, whereas for the low-frequency
peak χ = 16.4 kHz and η = 0.93.
Nitrogen-14 is something of a problem nucleus because there is no central transition
and quadrupole coupling constants can be significant (e.g., ~3 MHz)—sufficiently
high that MAS results in innumerable spinning sidebands and thus a loss of S/N.
Various schemes have been mooted to alleviate this situation, but there is no widely
used protocol yet.
FURTHER READING
“Quadrupole effects in NMR studies of solids”, M.H. Cohen & F. Reif, Solid State Physics , 5 (1957) 321–438.
“Quadrupole effects in solid-state NMR”, D. Freude & J. Haase, in NMR basic principles and progress , Eds. P.
Diehl, E. Fluck, H. Günther, R. Kosfeld & J. Seelig, Vol. 29 , Springer (1993) 1–90.
NOTES
1 The existence of a quadrupole moment corresponds to a nonspherical (but axially symmetrical) distribution of
positive charge within the nucleus. If the charge is greater along the spin axis than perpendicular to it, the
distribution is prolate. If the reverse is true, the quadrupole moment is oblate.
2 Alternatively, the symbol C can be used for the quadrupole coupling constant, though χ is recommended by
Q
IUPAC.
3 T.A. Wagler, W.A. Daunch, M. Panzner, W.J. Youngs & P.L. Rinaldi, J. Magn. Reson. 170 (2004) 336.
4 These ratios are correct only for nonselective pulses (see subsection 6.7.4).
5 Since the rotor and laboratory frames involve unique directions, only the unique angle between them is
involved in changing operators between these frames.
6 For a discussion of the conditions for selective excitation, see the article by Freude and Haase in Further
reading.
7 Actually, there are innumerable variations in the pulse sequences used for MQMAS, giving different
efficiencies and having various advantages and disadvantages. The subject is not for the faint-hearted! A
detailed evaluation is to be found in the article by Goldbourt and Madhu (see Further reading).
8 An alternative approach is to select only one pathway and generate a phase-modulated signal with respect to t
1 . See the review articles in Further reading for more information on this class of experiment.
9 In some articles, the scale in the 3Q dimension is divided by three to make it comparable to the scale of the
MAS dimension. However, this practice is not followed in comparable spin- 2D spectra and we believe it is
unhelpful.
10 D. Massiot, B. Touzo, D. Trumeau, J.P. Coutures, J. Virlet, P. Florian & P. J. Grandinetti, Solid State NMR 6
(1996) 73.
CHAPTER 7
7.1 INTRODUCTION
Figure 7.1. Examples of the timescales of motions in the solid state and the
NMR properties that can be used to explore t hem.
Section 7.3 shows how motion in the form of reorientation or, for example,
exchange of hydrogen, can affect the appearance of a spectrum.
Measurements designed to determine rates of motion are introduced there.
It is demonstrated how the anisotropic interactions important in solid-state
NMR (shielding/chemical shift anisotropy (CSA), dipolar coupling,
quadrupolar coupling) are modified by anisotropic motion and how
experiments can be used to probe details of this motion.
7.2 RELAXATION
Three modes of relaxation important to solid-state NMR were defined in
section 2.8. Spin–lattice (or longitudinal) relaxation, characterized by the
time constant T 1 , describes the process of restoring equilibrium to the z
component of the net magnetization for an ensemble of spins.1 Spin–spin
(or transverse) relaxation, with time constant T 2 , relates to the xy
component of the magnetization. Spin–lattice relaxation in the rotating
frame, with a time constant T 1 ρ , describes the return to equilibrium of
transverse magnetization in the presence of spin-locking from an RF
magnetic field.
Relaxation in NMR, unlike the case for most other spectroscopies, is non-
radiative . That is, the dominant cause of relaxation for the sample under
study is not by emission of electromagnetic photons. Instead, fluctuations
in the local field at the excited nucleus allow it to relax. Molecular motion
causes this fluctuation (see inset 7.1). In the solid state, the main
contributions to the local magnetic field arise from shielding anisotropy and
from dipolar coupling to the magnetic nuclei of surrounding atoms. Due to
its magnitude, the quadrupolar interaction, where present, often gives rise
to very efficient relaxation, though in this case it is a fluctuation in the local
electric field that is effective.
where τ c is called the correlation time and is the time constant that
characterizes the motion in question.3 Short correlation times imply fast
motion (as for liquids and solutions) and long ones slow motion (as for
rigid solids).
Figure 7.2. A log–log plot of the spectral density, given by equation 7.1, as
a function of correlation time for frequencies of: (a) 400 MHz, ( b) 800 MH
z , and (c) 62.5 kHz. For a given value of ν , J ( ν ) reaches a maximum
when 2π ντ c = 1 . For very slow motions (long correlation times),
indicated by the gray box, “classic” relaxation theory breaks down.
For a single, simple motion, the relaxation rates encountered in NMR can
be considered as being proportional to a linear combination of spectral
densities. For example, suppose the dominant relaxation mechanism is a
result of homonuclear dipolar coupling between spin- missing image file
nuclei, as might be the case for the protons in an organic solid. Under these
conditions, the relaxation times for a single spin species relate to the
spectral densities as follows:
Energy of Activation
Figure 7.4. An Arrhenius plot of the carbon spin–lattice relaxation rate for
the methyl group labeled (a) in ibuprofen. The gradient of the fitted line
gives an activation energy for methyl group rotation (assuming this is the
cause of the relaxation) of 7.9 kJ mol–1 .
Proton T 1 ρ values often provide extra information about the system. The
result of such a measurement on the same polyethylene sample is shown in
Figure 7.5 (b). This time, the signal decay is nonlinear and the relaxation
cannot be represented by one time constant. Such behavior is often an
indication of physical heterogeneity within a sample. For this polyethylene
sample, the T 1 relaxation is that expected of a homogeneous material but
the T 1 ρ behavior suggests heterogeneity. As discussed in inset 7.3, it is
likely that this sample is semicrystalline and that the average domain size is
between 5 and 50 nm. Analyzing relaxation data in terms of discrete
exponentials is one approach but it is not the only one, as shown in inset
7.4.
(b) The results from a T 1 ρ measurement on the same sample, showing the
signal S ( t ) as a function of spin-lock time (t ). The solid line is a least-
squares fit to the experimental data. The best fit is obtained from the sum of
three discrete exponential decays with time constants of T 1 ρ = 1.2, 4.5,
and 27.1 ms. Fitting curves like this to more than two components can be
contentious as the chemical/physical significance of the additional
component(s) may be debatable! Careful consideration of the error in each
value is essential. Both (a) and (b) were obtained with a solid echo (see
inset 3.5) and S ( t ) is the intensity of the echo maximum.
In a system with multiple relaxation time constants, each one will have a
population (the number of 1 H nuclei relaxing with that time constant)
associated with it. If there is no interaction between the domains to which
the time constants relate, then the populations will be proportional to the
number of 1 H atoms in the domains. However, if the domains interact so
that spin diffusion can occur between them, the populations of the faster
relaxing domain will be enhanced at the expense of that relaxing more
slowly and the measured populations will no longer reliably relate to the
relative numbers of 1 H atoms in each domain. In such cases, the lineshape
is a better indicator of the composition of a sample.
7.2.3 LINESHAPES & LINEWIDTHS FOR NON-SPINNING SAMPLES
Figure 7.6. (a) The static 1 H spectrum from polyethylene. (b) The
spectrum from the same sample obtained after spin locking the 1 H
magnetization for 18 ms. The part of the spectrum with the narrowest line
is largely lost and therefore must be associated with the shortest proton T 1
ρ value.
Figure 7.8. The decay of the proton spin-echo intensity for a non-spinning,
polycrystalline sample of glycine.
Interrupted Decoupling
Relaxation f ilters
Figure 7.11. The carbon-13 signal height for the crystalline component of a
polyethylene sample as a function of the mixing time following a
missing image file filter. A curve of this type can be used to deduce a
domain size for the amorphous component of the sample. Signal recovery
on this timescale and with this shape correlates with a 3D amorphous
domain size of the order of ~10 nm.
1 H–X correlation
The pulse sequences that are used to record the T 1 and T 1ρ relaxation
times discussed in this chapter have not yet been introduced. They are
grouped together in this section and are summarized in table 7.1.
Table Missing
Spin–lattice relaxation times
Figure 7.13 (c) and (d) incorporate CP steps. The pulse sequence in Figure
7.13 (c) is used to measure T 1 H indirectly through an X nucleus spectrum.
In a heterogeneous sample where T 1 H is not a sample-wide average (see
inset 7.3), it may be possible to relate T 1 H values to chemically
identifiable components within the sample (this information is not available
from a wideline 1 H measurement). The sequence shown in Figure 7.13 (d)
is for measuring T 1 X . Often, the delay τ has to extend to long times but
the advantages of CP discussed in section 3.4 apply here, making this
method practical for dilute and slowly relaxing nuclei. No 1 H decoupling
is applied during the variable delay to prevent damage to the probe. As a
consequence, relaxation behavior measured this way may be non-
exponential because there is the potential for cross relaxation involving
both 1 H and X nuclei and a simple analysis of the result may not be
possible.
In the basic measurement (Figure 7.14 (a)), spectra (or a single point from
the FID) are recorded as a function of the spin-lock time. The signal decay
is fitted to an equation of the form missing image file T 1ρ depends on
the RF field strength during the spin-lock (as well as on the sample) so it is
important to state this when quoting results.
Figure 7.14 (c) gives a way of measuring missing image file with the
advantages of the CP method. Proton decoupling is not applied during the
variable delay to prevent unwanted CP. Again, cross relaxation may
complicate the data analysis.
With all the spin-lock measurements, care must be taken not to damage the
probe with very long spin-lock pulses.
7.3 EXCHANGE
An exchange process can involve two or more sites (or environments). For
example, in an organometallic coordination polymer containing –C≡N–
Sn(CH3 )3 –N≡C–, there is rotation about the N–S n–N axis and the methyl
groups interchange between three equally populated environments. The
appearance of the carbon spectrum depends on the temperature of the
measurement: at low temperature three methyl signals are observed but at
high temperature a single signal is seen. For simplicity, exchange between
only two sites (A and B, say) is considered now:
If the forward ( k 1 ) and backward ( k 2 ) exchange rates are the same, then
A and B will be equally populated and they will contribute equally to the
NMR spectrum. If the exchange rates are not equal, then the system will
spend more time in one of the sites than the other (the lifetime, τ, of a
particular state is given by τ = 1/ k 1 or 1/ k 2 ). In this case, sites A and B
do not contribute equally to the spectrum.
Coalescence t emperature
where missing image file ν A and ν B being the frequencies of the two
lines related by exchange. For the case discussed here, the c and c′ lines
coalesce at approximately 25 ° C and Δ ν = 860 Hz, so missing image file
Figure 7.19. The difference in signal height as a function of mixing time for
the modified EXSY experiment on tropolone at 30 ° C. The time constant
of the decay is (ignoring relaxation) 1/2 k .
Energy of a ctivation
Deuterium
For spin- missing image file nuclei, the CSA can also be used to probe
motion. Two-dimensional experiments can distinguish between different
models of the reorientational process as the pattern of off-diagonal intensity
depends on the nature of that process. For 13 C, MAS is necessary for
sensitivity, so care is needed to ensure that the intensities of the off-
diagonal spinning sidebands relate to the exchange process (and are not
modulated by the sample spinning). Although 2D experiments can
distinguish between different motional models, acquiring a series of them
to extract exchange rates can be prohibitively time consuming. For this
reason, a number of experiments, which rely only on the measurement of
centerband intensities in 1D spectra, have been developed. One such is
CODEX (Centerband Only Detection of EXchange), and from two short
series of experiments, it is possible to obtain a correlation time and
information on the type of motion present (in principle for all resolved lines
in the spectrum) (see Figure 7.22 and Further reading).
In all the cases introduced so far in this chapter, the sample would be
described as a rigid solid. The general molecular motion in a “soft” solid
has a major impact on the NMR spectrum. For example, in a high-
resolution 13 C spectrum from such an organic material, the partial
averaging of the 13 C,1 H (and 1 H,1 H) dipolar coupling means that the
spectrum is to some extent self-decoupled and requires relatively little RF
decoupling power to reduce the linewidths to a small number of hertz (see
Figure 7.23 and inset 7.8). The CSA is fully averaged, so spinning
sidebands are not observed even at relatively low spin rates (<5 kHz).7
Even though such materials do not necessarily possess the long-range order
of a crystalline solid, the motional averaging of each environment (akin to
that in solution) means that the broad lines associated with a rigid
amorphous solid are not observed. In consequence, as illustrated in Figure
7.23 , highly resolved spectra can be obtained that begin to approach the
appearance of those from a solution.
Adamantane is a soft solid that gives very narrow lines at low (~2 kHz)
spin rates and low (~30 kHz ) decoupling fields (left). It is an excellent test
of shimming—recorded at 75.43 MH z, the full half-height linewidth here
is less than 2 Hz and satellites from 1 J CC can be observed. The spectrum
was obtained with CP.
7.3.6 INTERFERENCE
Molecular mobility can produce some unexpected effects in high-resolution
spectra when the rate of motion approaches that of MAS or the nutation
rate of 1 H decoupling. In both cases, the effect can be thought of as a
destructive interference so that the interactions which the spinning or
decoupling should be removing from spectra are being reintroduced.
Without going into the mathematics of the process here, it can be said that
the result is a broadening of the affected lines. This broadening will be
most extreme when the spinning or nutation and motion rates are matched,
but lines narrow again as they diverge. An example is given in Figure 7.24 .
Interference can also arise between the timing of some pulse sequences
and, for example, the sample spin rate as noted in section 5.5.1.
For the integration of any signal, it is important that the spectral baseline is
flat and has no offset (so the value of the integral truly represents the area
under the peak and not the shape of the baseline as well). If signals from
different samples are being compared, the signal intensity needs to be
corrected for the amount of sample in the rotor (so the sample mass should
be recorded). For a heterogeneous sample, it is important that the
components are uniformly distributed through the rotor (as the sample is
not uniformly excited/detected along the whole length of the rotor). When
signals overlap, spectral deconvolution (see inset 7.9) may be required. For
a resonance with spinning sidebands, the intensity of the latter should be
added to that of the centerband.
In Figure 7.25 the Lorentzian (gray line) is characterized by the broad base:
Figure 7.27 . Signal intensity vs. contact time for the CH labeled A
(circles) and carboxylic acid carbon (crosses) resonances in l -isoleucine.
Figure 7.28 . Schematic plot of signal intensity vs. contact time for a
heterogeneous system of two components with equal concentrations but
widely differing values of missing image file S A (max) and S B (max) is
the maximum amount of signal that can be observed for component A and
B, respectively.
The nature of the hyperfine coupling is complex and its magnitude (and
sign) varies from nucleus to nucleus. The coupling has essentially three
components: a Fermi contact interaction (due to unpaired electron density
at the nucleus), a pseudocontact term that is equivalent to the nuclear
dipole–dipole coupling, and a spin-orbit term (see Further reading for
details). Because the electron spins generally relax very quickly, the
multiplet resulting from the coupling is collapsed into a single line by
processes analogous to those described in section 7.3.13 However, the
position of this single line may be significantly different from that observed
from an analogous diamagnetic system. This paramagnetic shift arises
because the difference in energy of the electron spin states is not negligible
in comparison with kT. Consequently, the non-negligible population
difference for the electron spin states means that the population-weighted
average frequency is not at the center of the putative doublet (in the NMR
spectrum).
The hyperfine coupling is anisotropic, and the anisotropy can be very large
(thousands of parts per million even for 1 H spectra), so although it has the
usual p 2 (cosθ ) dependency, it can be difficult to handle with MAS. Thus
the MAS NMR spectra of paramagnetic materials are often largely
unresolved and contain extensive manifolds of spinning sidebands. As
shown in Figure 7.31 , however, very fast MAS (>25 kHz) may allow
useful resolution to be obtained.
FURTHER READING
“Multidimensional solid-state NMR and polymers” , K. Schmidt-Rohr &
H.W. Spiess, Academic Press (1994), ISBN 0 12 626630 1.
NOTES
2 Note that the symbol J does not relate to indirect coupling in this context!
5 Assuming the mixing time is much less than missing image file (in this
case).
9 The pair of spectra shown in Figure 7.23 (a) and (b) illustrate a similar
selectivity.
12 The nucleus at the site of a localized unpaired electron will have a very
large hyperfine coupling constant (potentially tens of MHz), resulting in
multiplets that are far too broad to be observed by conventional NMR.
13 The hyperfine coupling of the paramagnetic center itself is so large that
even the very quick electron spin relaxation is not efficient enough to
collapse the multiplet from the nucleus at the site of a localized unpaired
electron.
CHAPTER 8
8.1 INTRODUCTION
This chapter takes the reader through a variety of topics related to the
interpretation of spectra and the extraction of detailed information about the
NMR interactions. The first section describes the analysis of one-
dimensional spectra to obtain, for example, chemical shift information and
so does not involve special experimental techniques. However, obtaining
reliable estimates of other parameters, such as the strength of dipolar
couplings, may require specialist experiments and/or analysis, so most of
the rest of the chapter deals with somewhat more sophisticated aspects of
solid-state NMR, especially involving tensors and their evaluation. The
final section looks to the value of NMR in crystallography, where it
complements diffraction-based techniques (diffraction crystallography ).
8.2.1 GENERAL
Note that a common reference frame must be used for the tensors; here M
denotes a molecular frame of reference, rather than the unrelated principal
axis systems of the individual tensors. The parameters of the averaged
tensor can be fitted in the same way as “normal” unaveraged tensors,
although it is important to note that averaged tensors do not necessarily
have the same symmetry properties as the original tensors. For example,
averaged dipolar tensors may have non-zero asymmetries even though the
original dipolar tensors are axially symmetric by definition.
If the dynamic process is slow compared with the anisotropies (the slow
exchange limit), then tensor information can be determined independently
for the different sites (assuming that they can be resolved). In the
intermediate case, the form of the spectrum will be changed by the dynamic
process, with the exact shape depending on the tensors involved and the
nature of the dynamic process. Fitting such exchange-modulated spectra is
a powerful means of characterizing the dynamic process. The focus here is
on spectra that are unaffected by intermediate timescale dynamics.
If the sidebands of interest are well resolved from other peaks, then the
most straightforward approach is to determine the peak heights (or, better,
peak areas) of the observed sidebands. The intensity pattern of the spinning
sideband manifold is then fitted numerically to obtain the anisotropy
parameters. The Further reading has information about some of the various
software tools available.
The other inherent problem with dipolar couplings (in contrast to single-
spin interactions such as the shielding and quadrupole coupling) is the
promiscuous nature of through-space couplings; in a typical organic solid,
for example, a given 1 H spin will have significant dipolar couplings to a
number of nearby 1 H spins, resulting in a broad overall spectrum that
cannot be fitted in terms of individual couplings (see figure 2.5 ). As
mentioned above, selective labeling is often necessary to isolate an
individual pair of spins. Moreover, it cannot be assumed that an observed
dipolar coupling corresponds to an internuclear distance within the same
molecule—it could correspond to a short intermolecular distance.
Particularly in small-molecule systems and/or when accurate quantitative
results are required, it is common to recrystallize specifically labeled
molecules with an excess of the corresponding unlabeled molecule in order
to reduce the impact of intermolecular couplings.
Figure 8.4 illustrates two possible pulse sequences that can be used to
measure the heteronuclear dipolar couplings between a pair of spins, I and
S. The simplest approach (which works well if the coupling between I and
S is relatively strong) involves exploiting the CP experiment. If spin
diffusion among the I (and the S) spins is slow (compared with the
timescale for CP), then the CP build-up, obtained by measuring the
intensity of the S spin signal as a function of τ CP , is oscillatory. This is
illustrated in figure 8.5 for 1 H → 31 P CP in SnHPO4 . A strong
oscillation due to the 1 H,31 P dipolar coupling is observed. As discussed
in section 5.1.3, normal Hartmann–Hahn matching is inefficient in cases
where MAS is effectively suppressing 1 H spin diffusion; hence it is
necessary to match on a sideband, here n = 1. The “spectrum” associated
with this build-up curve then takes the form of a Pake-like pattern4 with the
separation of the horns given by missing image file .
Figure 8.4 . Two pulse sequences that can be used to measure heteronuclear
dipolar couplings: (a) cross-polarization (CP) signal acquired as a function
of CP contact time, τ CP , (b) a rotational-echo double-resonance REDOR
experiment, in which the S spin signal is measured after each spin echo,
with and without a train of rotor-synchronized 180° pulses applied to the I
spins. Note that phase cycling of the I spin pulses using so-called XY
schemes such as XY8 is widely used to minimize the effects of RF
imperfections.
Note that the chemical shift (including its anisotropy) is suppressed by the
spin-locking during the contact time and so the oscillation is purely a
function of the dipolar couplings. This experiment is then in effect a
separated local field experiment, since Fourier transformation with respect
to both τ CP and t 2 provides a spectrum with the dipolar coupling
information in one dimension and S spin chemical shifts in the other (the IS
coupling being suppressed in t 2 by the heteronuclear decoupling).
Additional weak IS couplings and spin diffusion will tend to broaden the
lines in the dipolar spectrum. As discussed in section 5.5.1, Lee-Goldburg
CP can be used to suppress the effects of spin diffusion to first order,
although this is unlikely to significantly sharpen the dipolar dimension for
powder samples.5
The second relevant pulse sequence shown in figure 8.4 (b) is the rotational
echo double resonance (REDOR) experiment, which uses recoupling (as
discussed in section 5.4.2) to reintroduce the effects of dipolar coupling that
are otherwise suppressed by MAS. In terms of the observed S spins, the
pulse sequence involves a spin echo, with the echo detected after an even
number of complete rotor periods (see Further reading). In the absence of
additional pulses on the I spins, the evolution due to the IS coupling
refocuses over a rotor period, while the chemical shift evolution is
refocused by the spin echo (see sections 4.6.1 and 4.4.2). Hence the only
evolution of the “reference” signal is due to homogeneous factors
(including relaxation) that are not refocused by the spin echo. If, however,
180° pulses are applied to the I spins twice per rotor period, then the
refocusing of the IS coupling is disrupted, and the observed signal is
reduced relative to the corresponding reference signal. Plotting the
(normalized) difference between the reference signal and the signal with
the REDOR pulses as a function of the dephasing time gives a build-up
curve which can be fitted to obtain the IS dipolar coupling. Figure 8.6
shows an example where NMR gives crystallographic information
complementing that from diffraction measurements. The aluminophosphate
sieve in question has cages containing a fluorine atom. Diffraction results
suggest this is centrally located in the cages, whereas the Al–F distance
determined by REDOR indicates that it is in a bonding position close to
one aluminum atom. The XRD and NMR results can be reconciled if the
fluorines are disordered over the multiple bonding sites in each cage. The
diffraction studies (erroneously) locate the fluorines at the cage centers.
Figure 8.6 . 27 Al{ 19 F} REDOR results for the Al signal (at 10 ppm) of
the aluminophosphate molecular sieve AlPO 4 -5. The open circles are the
experimental data. Calculated curves correspond to Al–F dipolar coupling
constants of (dotted line) 2.8 kHz ( r AlF = 0.219 nm) and (continuous line)
4.15 kHz ( r AlF = 0.192 nm). (This original figure was adapted for
publication in R.D. Gougeon et al., J. Phys. Chem. B 105 (2001) 12249.)
The REDOR experiment and its variants have been widely used to measure
dipolar couplings between a variety of nuclei. It can even be applied to
couplings between spin - and half-integer quadrupolar nuclei, with the
quadrupolar nucleus as the S spin, since refocusing the chemical shift
evolution at the midpoint of the recoupling sequence can be achieved
relatively cleanly via a single 180° pulse.
As the signal is only measured at (even) multiples of the rotor period, the
REDOR experiment is not suitable for strong dipolar couplings (of the
order of the spinning speed or greater) since not enough data points can be
obtained to characterize the dipolar oscillation. The CP-based experiment
of figure 8.4 (a) is better suited to these cases. In the other direction, the S
spin missing image file sets the ultimate limit on the smallest couplings
that can be measured. It is important to avoid overloading the probe when
using long recoupling times to measure weak couplings. For example, if 1
H decoupling is required, it will need to be applied continuously during
both recoupling and signal acquisition periods. It may be necessary to
truncate the acquisition (and sacrifice some resolution of the S spin
spectrum) to keep the duty cycle within acceptable limits.
In the case of 1 H spins in typical organic solids, there are multiple strong
homonuclear dipolar couplings between nearby H atoms, making it
impossible to isolate the effect of an individual coupling. The dynamics of
magnetization transfer between the spins are best described in terms of an
overall spin diffusion rather than the kind of oscillatory build-up curves
observed for isolated couplings seen in figure 8.5 . That said, the rate of
spin diffusion between a pair of spins will be directly related to the
coupling strength and can provide a qualitative indication of the proximity
of the spins involved.
where 2τ is the total spin-echo period (see sections 4.4.2 and 7.2.3), J is the
coupling constant, and missing image file is the time constant used to fit
the signal decay. Figure 8.9 shows an example of fitting J couplings
between pairs of 31 P nuclei. Couplings between different NMR nuclides
are fitted in the same way, except that the refocusing 180° is applied to both
nuclei involved to select for heteronuclear rather than homonuclear
couplings.
So far, it has mostly been assumed that only a single type of tensor affects
the spectrum, but clearly, in many (if not most) cases, several NMR
interactions contribute to defining it. Thus one must consider interplay
between tensors. Two such cases will be considered in some detail, one in
this section and a second one in section 8.6. Here the case of interplay
between the chemical shift and heteronuclear dipolar coupling will be
discussed. This will also involve indirect coupling. The situation can be
understood relatively simply because the tensors involved commute and
therefore affect the spectrum in an additive fashion. Spectra obtained at
more than one applied magnetic field can assist in understanding the
situation (since J and D are field independent to first order, whereas
chemical shifts scale linearly with the applied field), but the treatment of an
individual spectrum is considered here.
The simplest case, which illustrates the essential features, occurs for
isolated heteronuclear two-spin (AX) systems with missing image file in
which axial symmetry prevails so that asymmetry in σ A , σ X , and J AX
can be ignored. Moreover, all three tensors for, say, the A spins, will then
be coaxial (with the major principal axis along the internuclear A, X
distance, r AX ). The anisotropy in J AX has exactly the same form as the
dipolar coupling, D AX , and therefore an effective parameter
missing image file may be defined as:
where missing image file and missing image file are the A and X
Larmor frequencies in the absence of shielding, missing image file and
missing image file are the isotropic A and X shielding constants, ζ A and
ζ X are the shielding anisotropies (see inset 2.2) expressed as
missing image file and missing image file , respectively, J AX is the
isotropic (A, X) coupling constant, and θ is the angle between r AX and B 0
. The transitions of the A nucleus are therefore described by:
The equations above are given under the assumption that γ A and γ X are
both positive. Careful consideration needs to be given to signs if that is not
the case.
Equations 8.5 and 8.6 show that the A transitions form two subspectra
missing image file in one of which the shielding and dipolar tensors
reinforce one another, whereas in the other they partially cancel each other.
Thus, one static powder subspectrum is “stretched” whereas the other is
“squeezed”, as indicated schematically in figure 8.10 . This hypothetical
case is drawn for the case when J AX , D AX , and ζ A are all positive8
(involving missing image file ), with missing image file and
missing image file
missing image file
The situation is more complicated when the tensors involved are not
coaxial. This will be discussed qualitatively in section 8.7.
The effect on the energy levels is shown schematically in figure 8.13 . The
missing image file manifold of levels is affected by the second-order
terms in the opposite way to that of the missing image file levels. The
energy changes on the missing image file levels are equal whereas the
effect on the missing image file levels is opposite in sign and twice the
magnitude.
This means that the three allowed 13 C transitions are moved away from
the isotropic chemical shift, though the center of gravity remains
unchanged. The result is a 1:2 or 2:1 doublet, giving what is termed a
residual dipolar splitting . For the same reason that MAS does not eliminate
second-ord er effects on a quadrupolar spectrum, it cannot remove them
from the residual dipolar splitting of a spin - spectrum, though it does
cause significant averaging of the orientation-dependent factors. Therefore ,
the resonance frequencies depend on the angles between B 0 and the
dipolar and quadrupolar axes, so that the observed MAS spectrum for a
microcrystalline sample will be a powder pattern. Figure 8.14 illustrates the
spectrum for a 14 N, 15 N example, which clearly shows the splitting and
the powder-pa ttern nature of the two components. However, for the 13 C,
14 N case details of 13 C patterns are only usually seen at very low
magnetic fields; usually a broadened 2:1 or 1:2 doublet is recorded. For
such cases (under the assumptions already mentioned), the doublet splitting
is:
An equal term is required for the level m N = –1, whereas for the level m N
= 0, the analogous quantity is missing image file . The contribution to the
energy levels also involves the dipolar coupling constant and the polar
angles of the dipole direction in the magnetic field, but the net effect will be
proportional to a 2 and –2a 2 for the 14 N levels m N = ±1 and m N = 0
respectively. The sign of the second-order terms clearly depends on m C.
(However, terms in a 1 , involving the double step-up operator
missing image file and its partner, missing image file , are independent
of m C and therefore do not affect the I-spin transitions.)
Since missing image file is the Larmor frequency of the quadrupolar spin,
its value is known. Therefore, if χ is available (e.g., from NQR
measurements), D can be obtained (and therefore the internuclear distance
between the two spins derived). Conversely, if D is obtained from
diffraction measurements, χ can be estimated.
missing image file
The two preceding sections have discussed how NMR spectra are affected
when a given resonance is significantly influenced by more than one tensor
interaction, for example, a strong dipolar coupling combined with a large
shielding anisotropy. However, only cases with coaxial tensors were
examined. Frequently, this is not the situation in practice. Fitting NMR
parameters to spectra that are determined not only by the parameters of the
individual interactions but also by their relative orientations can be
difficult. However, the relative orientations of tensors form a potential
useful source of additional information, particularly in the context of
powder samples where it is impossible to establish the absolute orientations
of individual tensors. A wide variety of experiments has been proposed for
different combinations of dipolar coupling, quadrupolar coupling, shielding
tensors, etc., and this section simply summarizes the different broad
strategies in qualitative terms. Details can be found in the Further reading.
Static vs. MAS spectra can be useful when dealing with large quadrupole
couplings. The other anisotropic interactions are averaged out by
(sufficiently fast) MAS, leaving just a second-order quadrupole-broadened
lineshape (see section 6.4) that can provide accurate estimates for the
quadrupolar interactions, allowing the other interactions (and relative
orientations) to be determined from the static spectrum.
The multinuclear facet of NMR can allow deductions that go beyond the
determination of the crystallographic asymmetric unit to obtain molecular
symmetry information in the solid state. For example, the cyclic trimeric
organotin chalcogenides (Me2 SnX)3 , X = S or Se, give rise to two 119 Sn
signals (and two 77 Se signals in the selenide case) with intensity ratio 1:2,
showing that the molecule has symmetry. However, this could consist of
either a mirror plane or a twofold axis. In the former case, four 13 C signals
in intensity ratio 2:2:1:1 are predicted, whereas the existence of a twofold
axis would lead to only three signals (with equal intensities). The 13 C
spectrum of the sulfide shows unequivocally that the latter case is the
correct one. NMR can go further and determine the space group of a
crystalline solid. Thus, for the low-temperature phase of zirconium
phosphate, Zr2 P2 O7 , which was known to have a superstructure of the
cubic arrangement of the high-temperature phase (limiting the number of
compatible space groups), advanced NMR techniques (such as those
discussed in section 5.4.4) applied to the 31 P spectrum (figure 8.19 )
shows the existence of 27 peaks arising from 14 diphosphate ions, one (and
one only) of which contained phosphorus atoms related by symmetry. This
information established that the space group is Pbca , enabling the powder
diffraction pattern to be solved to give the full crystal structure.
This synergy between NMR and diffraction can be taken further. For
instance, solving structures from powder diffraction data by simulated
annealing procedures can be done using computer programs that introduce
constraints based on NMR information (interatomic distances from dipolar
coupling constants or knowledge of intermolecular hydrogen bonding from
chemical shifts). In the future, it may prove to be possible to
simultaneously refine structures to fit diffraction information and NMR
spectra. Already, NMR can be used to refine structures obtained from
diffraction data, as in the case of zeolite ZSM-12, where a relatively poor
structure obtained by powder X-ray diffraction was improved to give good
agreement between the principal components of the 29 Si shielding tensors
while retaining acceptable computed diffraction patterns. It has also been
shown that 13 C tensor components for crystalline naphthalene reveal
subtle distinctions between the geometry at different carbon atoms that are
consistent with the diffraction-determined space group, whereas the
diffraction-determined geometry gives no such distinctions outside
experimental error.
Figure 8.21 . View along the axis of the aromatic group in phenobarbital
(scheme 8.2) form III to illustrate how the nitrogen atoms N1 and N3 are
differentiated in the crystal structure.
“NMR crystallography” , Eds. R.K. Harris, R.E. Wasylishen & M.J. Duer,
John Wiley & Sons Ltd. (2009), ISBN 978-0-470-69961-4.
“Dipolar coupling: Its measurement and uses”, M.J. Duer, Chapter 3 , pp.
111–178 (DOI: 10.1002/9780470999394.ch3) and “Molecular structure
determination: Applications to biology”, O.N. Antzutkin, Chapter 7 , pp.
280–390 (DOI: 10.1002/9780470999394.ch7) of “Solid-state NMR
spectroscopy: Principles and applications” , Ed. M.J. Duer, Blackwell
(2002), ISBN 0.632-05351-8.
“Measuring heteronuclear dipolar couplings for I = 1/2, S > 1/2 spin pairs
by REDOR and REAPDOR NMR”, T. Gullion & A.J. Vega, Prog. NMR
Spectry ., 47 (2005), 123–136. DOI: 10.1016/j.pnmrs.2005.08.004.
NOTES
3 Note that the spectra would also be dependent on the relative orientations
of the tensors involved, which highlights the difficulties of quantifying
spectra when multiple interactions are present. Section 8.7 discusses the
question of relative tensor orientations in more detail.
4 The pattern is not the same as the normal Pake pattern illustrated in figure
2.4 for a heteronuclear spin pair (Δ = 2D ) due to the spin-lock conditions
and the n = ± 1 sideband matching. The splitting is even smaller (Δ = D/ 2)
for the n = ± 2 sidebands, making them a less desirable option.
Table Missing
Table Missing
APPENDIX B
LIOUVILLE S PACE , R
ELAXATION & E XCHANGE
C.3
C.4
C.5
that is, the coherence oscillates with a frequency Δ . Note how the
evolution of the density matrix can be read directly from equation C.1: the
populations, and , do not evolve, while the and
coherences evolve at frequencies of Δ and –Δ , respectively.
The Liouville space treatment is a rather extravagant way of describing
free precession. Consider, however, the Liouvillian and propagator:
C.6
hence
C.7
C.8
C.9
where is a relaxation matrix as in equation C.6. Note how will give rise
In problems involving slow exchange (i.e., dynamics that are much slower
than the scale of the NMR interactions), it is often possible to avoid
Liouville space treatments and simply consider the exchange of z
magnetization during magnetization transfer or EXchange SpectroscopY-
like experiments. If we consider, for example, a system in which z
magnetization is exchanging between three sites, A, B, and C, with the
following rates for individual transfers
C.10
C.11
where M is the vector containing the site magnetizations, M k , and K is the
exchange matrix . Note how the structure of K indicates which sites are
exchanging, and how mass balance means that each column must sum to
zero. Note also that K is not generally symmetrical, and care must be taken
not to confuse rows and columns, forwards vs . backwards exchange rates,
etc.
Equation C.11 has exactly the same form as equation C.2 and again has
the solution
C.12
C.13
C.14
C.15
C.16
C.17
C.18
FURTHER READING
Chapter 2 of “Principles of nuclear magnetic resonance in one and two dimensions” , R.R. Ernst, G.
Bodenhausen & A. Wokaun, Oxford University Press, (1990), ISBN 0 19 855647 0.
NOTES
INTRODUCTION TO SOLID -
STATE NMR SIMULATION
The first step in any simulation is specifying the set of nuclear spins and
their NMR interactions needed to describe the problem at hand. Calculation
times typically increase by almost an order of magnitude for each additional
spin, and using the smallest spin system that reproduces the experimental
results is vital to obtaining results in a reasonable length of time (as well as
simplifying the resulting spectra). Fortunately it is rarely necessary to
specify more than a couple of spins. For instance, the spectra of
quadrupolar nuclei are dominated by the quadrupolar interaction and the
coupling to other nuclear spins can often be ignored. Hence, a simple one-
spin simulation involving just the quadrupolar interaction (and possibly the
chemical shift anisotropy (CSA), depending on its size) should be sufficient
to model the spectrum. Similarly, when fitting the 13 C CSA pattern, it is
not necessary to consider coupling to surrounding protons. Although
residual dipolar coupling will broaden the 13 C resonances, this does not
need to be accounted for explicitly. Simulations involving abundant spins
are considerably more challenging; calculating the lineshape associated
with a set of coupled spins necessarily involves using a sufficiently large
number of spins (typically about 10) in order to reproduce experimental
results.
As an example, the spin system for a two-spin heteronuclear problem
may be expressed (SIMPSON and pNMRsim) by
spinsys {
nuclei 13C 1H
dipole 1 2 -20e3 0 0 0
shift 1 10p 50p 0.5 0 30 0
}
where dipole specifies the coupling between the spins to be 20 kHz, with its
principal axis system aligned with the molecular axis system, that is with
Euler angles Ω MP = (0, 0, 0). The shift line specifies the isotropic chemical
shift to be 10 ppm and the CSA to be 50 ppm with an asymmetry parameter
of 0.5 and with a PAS tilted by 30° away from the molecular frame. Note
how the anisotropy information, including relative orientations, must be
fully specified for the calculation to be properly defined, which is another
motivation for using the smallest possible spin system!
D.2 SPECIFYING THE POWDER SAMPLING
crystal_file zcw132
gamma_angles 8
channels 1H
…
start_operator 13C:x
detect_operator 13C:+
…
pulse 5 100e3 x
pulse 5 100e3 -x
store XiX
…
pulseq {
acq 0 XiX
}
FURTHER READING
“Numerical simulation of solid-state NMR experiments”, P. Hodgkinson & L. Emsley, Prog. Nucl.
Magn. Reson. Spectrosc ., 36 (2000), 201. DOI: 10.1016/S0079-6565(99)00019-9.
“Computer simulations in solid-state NMR”, Mattias Edén, Concepts Magn. Reson ., 17A (2003),
117; 18A (2003), 1; 18A (2003), 24. DOI: 10.1002/cmr.a.10061; 10.1002/cmr.a.10064;
10.1002/cmr.a.10065.
SIMULATION PROGRAMS
Accumulation time, 8
Acetonitrile, 240
Acquisition
window, 134
Activation
enthalpy, 182
entropy, 182
Alanine, 3, 161
Aluminosilicate, 156
preferred orientation, 11
Amplitude
of motion, 201
Analysis
Anisotropic
diffusion, 203–204
Arcing, 53–54
B
Barbital, 6, 7
Background signal, 69
for quadrupolar coupling, 32, 152–154, 166–167, 167, 199, 201, 217, 238
for shielding, 23
Centerband
Central transition
Chemical shift, 4, 21, 72, 81, 85, 88, 158, 177, 212, 215, 228, 241. See also
Shielding
anisotropy (CSA), 99, 102, 110, 132–133, 159, 201, 221, 224, 227, 259–
260. See also Shielding anisotropy
computation, 244
referencing, 54
Computer-simulation/computation
Coherence
selection, 63
zero-quantum (ZQ), 93
Correlation experiments/spectra
EXSY, 199–202
in general, 120–122
heteronuclear. See HETCOR, WISE
Cortisone, 112
Coupling
homonuclear dipolar, 26, 35, 102, 107, 116, 180, 186, 196, 201, 216, 225–
228
indirect (J), 20–21, 24, 117–120, 126, 130, 228, 231, 239
CPMG, 188
auto-correlation, 129–131
Cross polarization (CP), 15, 58–66 , 95, 113–115 , 121, 123–125, 158, 191,
208, 222–223. See also Lee-Goldburg
1 H to 19 F, 67
contact pulse, 59
ramped, 114–115
spin-lock, 59–60
Crystal
lattice, 8
Crystalline
state/form, 4–6
BLEW-12, 119
BR-24, 134
SPINAL64, 112
reduced, 84
Digital resolution, 67
Dilute nuclides, 15 , 58
Dipolar dephasing, 116, 119, 136, 189. See also Decoupling, interrupted
Direct dimension, 120–122 , 129, 134, 164
Domain, 11, 183, 184, 192, 209. See also Heterogeneous materials
DUMBO, 134
Echo
shifted, 169
Enantiotropic system, 12
13 C, 14 N system, 234
Equivalence, 4, 131
ERETIC, 210
of the density matrix, 84, 88, 91, 101, 103, 105–106, 255, 263
Exchange
CODEX, 202–203
1-Fluorononane, 13
Free-induction decay (FID), 8, 36, 41 , 44–50, 53, 67, 145, 157, 162, 172,
188, 195, 262
Goniometer, 220
Hamiltonian, 71–76
Zeeman, 73 , 77–78, 80
theoretical treatment, 96
Hg(SCN)2 , 51
High-field approximation, 73 , 97
HMQC, 127
Host, guest systems, 13–14. See also Finasteride; Urea inclusion systems
Hydrogen bonding, 13
Ibuprofen, 182
Incommensurate structures, 10
Indirect coupling. See Coupling
Indomethacin, 27–28
frame, 79 , 103–104
internal, 2, 18
Interferogram, 41
Internuclear distance, 25
Inversion-recovery, 194–195
Isoleucine, 208
Isotropic average/component
of adamantane 8, 205
Knight shift, 23
Lee-Goldburg, 68
decoupling, 134
inhomogeneous, 115–117
intrinsic, 116
electric, 179
magnetic, 178
Low-gamma nuclei, 52
definition of, 33
CRAMPS, 133
Magnetic moment, 17
TEDOR, 224
algebra, 78, 86
exchange, 256
representations of operators/Hamiltonians, 74–75
relaxation, 255
representation of tensors, 18
MAT, 220
MELODRAMA, 227
Methanol, 238
diffusive, 203–204
reduced dipolar coupling, impact on spectra of, 53, 187, 190, 224
reorientation, 201–203
MQMAS, 163–169
Naphthalene, 242
Nutation
curve, 42 , 168
spectrum, 159
T 1ρ relaxation, 180
Octosilicate, 124
Off-resonance, 44 , 110. See also Decoupling, Lee-Goldburg
detection, 132
Organometallics, 212
Oversampling, 49
Oxygen-17, 17 O, 161
PASS, 221
PDLF, 223
Phase modulation,
Phenol, 9
PISEMA, 223
PMLG, 135
Polymorphism, 7, 12 , 167
Population-weighted average
frequency, 212
Powder pattern, 3, 14
Principal axis system (PAS), 19 , 22, 78, 99, 100, 147–148, 216, 260
Probe ringing, 42
Pseudopotentials, 224
Pulse
breakthrough, 50
flip-back, 65
reconversion, 164
CPMG, 188
EXSY, 199
flip-back, 65
FSLG, 135
HETCOR, 124
INADEQUATE, 128
MQMAS, 164
REDOR, 223
RFDR, 127
STMAS, 172
Torchia, 194
TPPM, 115
WISE, 192
QCPMG, 158
Quadrupolar nuclei,
Quadrature detection, 61
in 2D NMR, 122
Quantitation
MELODRAMA, 227
REDOR, 223–225
intensity, 208
Relative intensity. See Quantitation
filtering, 191
T 2 *, 188–192
T 2 ′ , 119, 188
Relaxometry, 38
Resolution, 3, 14, 67
digital, 67
RF inhomogeneity, 43 , 113
Sensitivity of NMR, 18
Shielding, 2, 4, 19–23 , 76
Shielding anisotropy, 14, 21–23 , 33–35, 37, 159, 216, 218, 220–222, 230,
237
Signal-to-noise ratio, 45
Silk, 11
Sildenafil citrate, 14
Simulation/Simulated spectra, 35, 162, 201, 207, 217, 237, 241, 243, 259–
263
Spikelets, 158
Spin-temperature inversion, 62
Static samples, 14. See also Powder pattern, Single crystal (NMR of)
STMAS, 172–173
Sulfathiazole, 13
T 2 , T 2 ′ , T 2 *. See Relaxation
TEDOR, 224
Tensor, 18–21 , 78
dipolar coupling, 20–21, 24–25 , 78, 239
interplay, 228–233
principal components, 19
shielding, 18–23, 76
spherical, 97–99
Time domain, 41
Trimethylphosphine sulphide, 37
Tropolone, 199
Unit cell, 4
WISE, 192
z-filter, 166
Zero-filling, 47, 67
____________________________________________