Akitt, J. W. - Mann, Brian E - NMR and Chemistry - An Introduction To Modern NMR Spectroscopy, Fourth Edition-CRC Press (2017)
Akitt, J. W. - Mann, Brian E - NMR and Chemistry - An Introduction To Modern NMR Spectroscopy, Fourth Edition-CRC Press (2017)
www.pdfgrip.com
www.pdfgrip.com
NMR and Chemistry
An introduction to modern
NMR spectroscopy
Fourth Edition
Dr J.W.Akitt
Formerly Senior Research Officer at the
University of Leeds
A CRC title, part of the Taylor & Francis imprint, a member of the
Taylor & Francis Group, the academic division of T&F Informa pic.
www.pdfgrip.com
© 2000 J. W. Akitt and B. E. Mann
10 9 8 7 6 5 4
A catalogue record for this book is available from the British Library
ISBN 0 7487 4344 8
Every effort has been made to contact copyright holders of any material reproduced
in this book and we apologise if any have been overlooked.
www.pdfgrip.com
Contents
Abbreviations ix
Preface to the fourth edition xv
www.pdfgrip.com
vi Contents
www.pdfgrip.com
Contents vii
Bibliography 385
Answers to Questions 391
Index 397
www.pdfgrip.com
www.pdfgrip.com
Abbreviations
www.pdfgrip.com
x Abbreviations
www.pdfgrip.com
Abbreviations xi
E0 an energy barrier
Ea Arrhenius activation energy
Ez electric field along a bond
F Nyquist frequency
G field gradient
Gx, Gy, Gz field gradients in the jc, y, and z directions
h reduced Planck's constant, h/2n
I moment of inertia
nuclear spin quantum number
n
j coupling constant through n bonds (in Hz)
n
K reduced coupling constant 47i2/MyAyB, through n bonds
k Boltzmann's constant, 1.3807 x 10~23 J/K
rate constant
K(v) field intensity at a frequency v
m angular momentum quantum number
Mx net magnetization along the x axis
Mxy net magnetization in the xy plane
My net magnetization along the y axis
Mz net magnetization along the z axis
P power
pressure
PA> PE relative populations of sites A and B
Q electric quadrupole moment
q{ electric charge
R relaxation rate, s~!
the gas constant, 8.3145 J Kr1 moH
r a distance
Rl spin-lattice relaxation rate, s'1
^IDD dipole-dipole spin-lattice relaxation rate, s'1
RIQ quadrupolar spin-lattice relaxation rate, s-1
/?1SR spin-rotation spin lattice relaxation rate, s"1
R2 spin-spin relaxation rate, s"1
^200 dipole-dipole spin-spin relaxation rate, s'1
^?2Q quadrupolar spin-spin relaxation rate, s"1
S electron spin quantum number
nuclear spin quantum number of nucleus 5,
T relaxation time, s
T time, s
Tl spin-lattice relaxation time, s
7\CSA chemical shift anisotropy spin-lattice relaxation time, s
riDD dipole-dipole spin-lattice relaxation time, s
7\e electron spin-lattice relaxation time, s
rlQ quadrupole spin-lattice relaxation time, s
T2Q quadrupole spin-spin relaxation time, s
risc scalar coupling spin-lattice relaxation time, s
r2SC scalar coupling spin-spin relaxation time, s
riSR spin-rotation spin-lattice relaxation time, s
T2 spin-spin relaxation time, s
www.pdfgrip.com
xii Abbreviations
www.pdfgrip.com
Abbreviations xiii
www.pdfgrip.com
www.pdfgrip.com
Preface to the
Fourth Edition
It is some six years since J.W. Akitt wrote the preface to the third
edition and now, in retirement, he welcomes the cooperation of an
old colleague, Professor Brian Mann of the University of Sheffield, to
update the text. The first edition appeared in 1972, over 26 years ago,
and in this time NMR spectroscopy has seen immense changes and
developments. The pace of technical change has perhaps slowed over
the six years since the third edition appeared but the number of appli-
cations continues to increase and it has become time to make big
changes in the presentation of the book while keeping the general
layout of the subjects. High-field spectrometers have become common-
place and their operation is carried out via comprehensive computer
control so that they are easy to use and are now operated even by
undergraduate students as a teaching facility. This has led us to intro-
duce rather more description of the workings and operation of spec-
trometers to help such debutant users to understand better what their
commands actually cause to happen.
One topic that has been finally abandoned in this edition is the old
method of continuous wave NMR spectroscopy. Also, to a large extent,
the old CW NMR spectra, complete with wiggle beats have been
replaced by Fourier transform ones. However, we should remember
that the time-shared lock systems used on the FT equipment are best
understood in CW terms.
Now that multinuclear NMR spectroscopy has become routine with
a wide range of nuclei being easily observable, many examples of their
use have been included, ranging from sensitive nuclei such as 31P to
very insensitive nuclei such as 187Os. The discussion of multipulse and
two-dimensional NMR spectroscopy has also been enlarged to reflect
their increasing importance. Many such experiments have become
routine, with fully automated instruments running them.
Examples and many new problems, mainly from the recent inor-
ganic/organometallic chemistry literature, support the text throughout.
Brief answers to all problems are provided in the text with fuller
answers on the Thornes website at http://www.thorneseducation.com.
We are indebted to many people for their comments and assistance,
especially to Professor K. Elsevier for commenting on early drafts of
parts of the manuscript, Bruker Spectrospin Ltd for insights into
present day electronics, and for permission to reproduce diagrams from
www.pdfgrip.com
xvi Preface
www.pdfgrip.com
The theory of nuclear
magnetization 1
1.1 THE PROPERTIES OF THE NUCLEUS OF AN ATOM
www.pdfgrip.com
2 The theory of nuclear magnetization
see that the sign of 7 influences both spin-spin coupling and the way
energy is exchanged between spins.
If / > 1/2 the nucleus possesses in addition an electric quadrupole
moment, Q. This means that the distribution of charge in the nucleus
is non-spherical and that it can interact with electric field gradients
arising from the electric charge distribution in the molecule. This inter-
action provides a means by which the nucleus can exchange energy
with the molecule in which it is situated and affects certain NMR
spectra profoundly.
Some nuclei have 1 = 0. Important examples are the major iso-
topes 12C and 16O, which are both magnetically inactive - a fact that
leads to considerable simplification of the NMR spectra of organic
molecules. Such nuclei are, of course, free to rotate in the classical
sense, but this must not be confused with the concept of quantum-
mechanical 'spin'. The nucleons, i.e. the particles such as neutrons and
protons which make up the nucleus, possess intrinsic spin in the same
way as do electrons in atoms. Nucleons of opposite spin can pair,
just as do electrons, though they can only pair with nucleons of the
same kind. Thus in a nucleus with even numbers of both protons and
neutrons all the spins are paired and / = 0. If there are odd numbers
of either or of both, then the spin is non-zero, though its actual value
depends upon orbital-type internucleon interactions. Thus we build up
a picture of the nucleus in which the different resolved angular
momenta in a magnetic field imply different nucleon arrangements
within the nucleus, the number of spin states depending upon the
number of possible arrangements. If we add to this picture the concept
that s bonding electrons have finite charge density within the nucleus
and become partly nucleon in character, then we can see that these
spin states might be perturbed by the hybridization of the bonding elec-
trons and that information derived from the nuclear states might lead
indirectly to information about the electronic system and its chemistry.
The most important properties of the elements relevant to NMR
spectroscopy are listed in Table 1.1. This gives the atomic weight of
the nuclear isotope of the element listed, and where there is more
than one magnetically active isotope this is indicated by an atomic
weight in parentheses. The isotope listed is the one most usually used,
though the unlisted ones are in some cases equally usable. The next
column gives the spin quantum number, /, followed by the natural
abundance of the isotope. The receptivity or natural signal strength of
the nucleus is given relative to that of 13C, which itself gives a fairly
weak signal, and this figure is made up of the intrinsic sensitivity of
the nucleus (high if the magnetic moment is high) weighted by its
natural abundance. Some elements are used in enriched forms (particu-
larly 2H and 17O), when the receptivity is, of course, substantially
higher. The data are completed by the quadrupole moment (where / >
1/2) and the resonant frequency in a particular magnetic field. This
can be determined to a much higher precision than shown, for a reso-
nance in an individual compound, and it should be remembered that
www.pdfgrip.com
The properties of the nucleus of an atom 3
www.pdfgrip.com
4 The theory of nuclear magnetization
each nucleus will have a range of frequencies due to the chemical shift
effect. The resonance frequency is proportional to the magnetogyric
ratio.
The first point to note about this list is that almost all the elements
are represented, the only missing stable ones being argon and cerium,
and that, in principle, virtually the whole of the Periodic Table can be
studied by NMR. Indeed, with modern instrumentation, this is now
realizable, though there are some cases, notably nuclei with very high
quadrupole moments, where study in the liquid state is not rewarding.
The usefulness of a nucleus to the NMR spectroscopist depends in the
first place upon the chemical importance of the atom it characterizes
and then upon its receptivity. Thus the extreme importance of carbon
www.pdfgrip.com
The nucleus in a magnetic field 5
www.pdfgrip.com
6 The theory of nuclear magnetization
(a)
Figure 1.1 (a) The nuclear spin energy for a single nucleus with 1=1 (e.g. 14N) plotted as a function of
magnetic field B0. The two degenerate transitions are shown for a particular value of B0 (b) The align-
ment of the nuclear vectors relative to BQ that correspond to each value of m. The vector length is
ftV/(7 + 1) and its z component is hm, whence cos 9 = mN 1(1 + 1).
hv = A£
where AE1 is the energy separation. For NMR
hv = fjjJo//
In this case, the transition for any nuclear isotope occurs at a single
frequency since all the energy separations are equal and transitions
are only allowed between adjacent levels (i.e. the selection rule Am =
± 1 operates). The frequency relation is normally written in terms of
the magnetogyric ratio (1.1), giving
v = 7^0/211 (1.2)
Thus the nucleus can interact with radiation whose frequency
depends only on the applied magnetic field and the nature of the
nucleus. Magnetic resonance spectroscopy is unique in that we can
choose our spectrometer frequency at will, though within the limita-
tion of available magnetic fields. The values of y are such that for
practical magnets the frequency for nuclei lies in the frequency range
between a present maximum of 800 megahertz (MHz) and a minimum
of a few kilohertz (kHz) using the earth's magnetic field.
www.pdfgrip.com
The source of the NMR signal 7
www.pdfgrip.com
8 The theory of nuclear magnetization
x My=Mv=0
(a) (b)
Figure 1.2 Freely precessing nuclei in a magnetic field B0. (a) Larmor precession of a single nucleus,
(b) The excess low-energy nuclei in a sample. The nuclear vectors can be regarded as being spread evenly
over a conical surface. They arise from different atoms but are drawn with the same origin.
www.pdfgrip.com
The source of the NMR signal 9
Figure 1.3 The RF current in the coil in Fig. 1.4 produces an oscillating magnetic field along the y axis.
This can be equally well regarded as being composed of two rotating magnetic vectors rotating in oppo-
site directions, whose resultant is the oscillating magnetic field. One of these vectors will rotate in the
same sense as the nuclear precession and is conventionally called Bv
www.pdfgrip.com
10 The theory of nuclear magnetization
coil
M y =0
Figure 1.4 If a rotating magnetic vector Bl with the same angular velocity as
the nuclei is now added to the system, the nuclei will also tend to precess
around Bl9 and this causes the cone of vectors to tip and wobble at the nuclear
precession frequency. The resulting rotating vector Mxy in the xy plane can
induce a current in the coils that are wound beside the sample.
www.pdfgrip.com
The source of the NMR signal 11
Figure 1.5 A sufficiently long or powerful B{ along the x' axis will turn the magnetization into the xy
(x'y') plane and the whole of the nuclear magnetization lies along the / axis. This is in the rotating frame
and My> rotates within the static laboratory frame, contributing to the signal picked up by the coil in Fig.
1.4 and the output is at a maximum. Bl is thus applied in the form of a pulse, and a pulse that has the
effect illustrated is known as a 90° pulse. In the rotating frame, Bl remains pointing in a fixed direction
and the magnetization rotates about this direction. In the laboratory frame, Bl rotates in the xy plane and
the magnetization follows a spiral path away from the BQ axis during the pulse.
all nuclei with precession frequencies situated within this band. The
strength of B\ at frequencies around the Bl frequency varies as sin(x)/x
(Fig. 1.6), though the intensity distribution is reasonably constant near
Br In fact, all nuclei within ±l/4PWHz of the spectrometer frequency
(and therefore of the Bl frequency) are almost equally affected. PW
is the pulse width or duration. There are, in contrast, null points at
±n/PW Hz where the nuclei are not perturbed. The negative intensi-
ties at higher frequency separation indicate a 180° phase change where
the nuclei would swing in the opposite direction.
Evidently, we are interested only in those nuclei in the central region
of the frequency distribution. For these, the magnetization in the labo-
ratory xy plane, M precesses about B0 with the Larmor frequency
and continues to do so after B1 has been switched off. This rotating
www.pdfgrip.com
12 The theory of nuclear magnetization
Figure 1.7 A free induction decay, FID, from a single type of nucleus. This signal contains the same infor-
mation as a conventional NMR spectrum. The frequency of the signal relative to the spectrometer frequency
is given by the frequency of the cosine wave. The intensity is given by the intensity of the cosine wave at
its beginning, and the linewidth is given by the rate of decay of the cosine wave.
www.pdfgrip.com
A basic NMR spectrometer 13
-* 1 Magnet I
M vy
Power rr—^—i
amplifier | lReceiverl
I Printer [/
www.pdfgrip.com
14 The theory of nuclear magnetization
1.5 QUESTIONS
www.pdfgrip.com
Questions 15
1.7. Calculate the angle at which a single proton will precess about a
magnetic field.
1.8. What is the receptivity of the lithium nucleus, 6Li, after enrich-
ment to 100%?
1.9. The nucleus, 1H, in water resonates at 400 MHz in a magnetic
field of 9.39 T. The earth's magnetic field is 0.000 05 T. What is
the !H precession frequency in the earth's magnetic field and
what is the excess population of nuclei in the lower energy state
in this field at 300 K?
1.10. A resonance in a given sample and spectrometer has a frequency
of 100 000 000 Hz. A continuous monochromatic Bl of frequency
100 000 002 Hz is applied in the xy plane to the sample. Because
Bl and nuclear frequencies are not the same, Mz will tend to
rotate around the z axis and never leave it to reach the xy plane.
What will be the frequency of rotation of Mz around the z axis
in the rotating frame?
www.pdfgrip.com
www.pdfgrip.com
The magnetic field at
the nucleus: nuclear
screening and the
2
chemical shift
So far we have shown that a single isotope gives rise to a single nuclear
magnetic resonance in an applied magnetic field. This really would be
of limited interest to the chemist except for the fact that the magnetic
field at the nucleus is never equal to the applied field, but depends in
many ways upon the structure of the molecule in which the atom
carrying the nucleus resides.
The most obvious source of perturbation of the field is that which
occurs directly through space due to nuclear magnets in other atoms
in the molecule. If such nuclei have high magnetic moments, which
means generally !H or 19F, then in the solid state this interaction results
in considerable broadening of the resonance, which obscures much
information. However, in the liquid state, where the molecules rotate
rapidly and randomly, the direct nuclear fields fluctuate wildly in both
intensity and direction and have an average value that is exactly zero.
The resonances are thus narrow and may show much structure. Thus
spectroscopy of the liquid state has been a major preoccupation of
chemists.
Since the magnetic nuclei do not directly perturb the field at the
nucleus, we have therefore to consider the effect that the electrons in
the molecule may have. We will concern ourselves only with diamag-
netic molecules at this stage and will defer till later discussion of para-
magnetic molecules possessing an unpaired electron. When an atom
or molecule is placed in a magnetic field, the field induces motion of
the electron cloud such that a secondary magnetic field is set up. We
can think of the electrons as forming a current loop as in Fig. 2.1
centred on a positively charged atomic nucleus. The secondary field
produced by this current loop opposes the main field at the nucleus
and so reduces the nuclear frequency. The magnitude of the electronic
current is proportional to BQ and we say that the nucleus is screened
www.pdfgrip.com
18 The magnetic field at the nucleus
Figure 2.1 The motion of the electronic cloud E around the nucleus N gives
rise to a magnetic field, shown by dashed lines, which opposes BQ at the
nucleus.
(or shielded) from the applied field by its electrons. This concept is
introduced into equation (1.2) relating field and nuclear frequency by
the inclusion of a screening constant a
»d-,) (2.1)
ZTT
a is a small dimensionless quantity and is usually recorded in parts
per million (ppm). The screening effect is related to the mechanism
that gives rise to the diamagnetism of materials and is called diamag-
netic screening.
The magnitude of the effect also depends upon the density of elec-
trons in the current loop. This is a maximum for a free atom where
the electrons can circulate freely, but in a molecule the free circula-
tion around an individual nucleus is hindered by the bonding and by
the presence of other positive centres, so that the screening is reduced
and the nuclear frequency increased. Since this mechanism reduces
the diamagnetic screening it is known as a paramagnetic effect. This
is unfortunately a misleading term and it must be emphasized that it
does not imply the presence of unpaired electrons. As used here, it
merely indicates that there are two contributions to a, the diamag-
netic term crd and the paramagnetic term <rp and that these are opposite
in sign. Thus
o- - crd + ap (2.2)
Since the magnitude of <rd depends upon the density of circulating
electrons, it is common to find in the literature discussion of the effect
of inductive electron drifts on the screening of nuclei. The screening
of protons in organic molecules, for instance, depends markedly on
the substituents, and good linear correlations have been found between
screening constants and substituent electronegativity, thus supporting
www.pdfgrip.com
Effects due to the molecule 19
www.pdfgrip.com
20 The magnetic field at the nucleus
90001- - v V •y.
8000-
7000
6000
ppm 5000
4000 -
'-
3000
2000
1000H
0 Hi Li , 1 , ,1 t T. iY 1 A
B N F Nal Al p Cll Kl fcdGd As[BrlRbl Ag InSb \ Cd "Adrri
HeBe (: c) NeMg S >i J5 ArCa ZnG eS.e Kr Sr CdSnTeXeBa H gPb
Figure 2.2 Ranges of screening constants for nuclei of main-group and post-transition elements. It should
be remembered that, while the ranges shown reflect the periodicity of the (1/r3) term, they also are influ-
enced by how many compounds of a given element have been measured and, indeed, by the extent of its
chemistry. (After Jameson and Mason (1987) Multinuclear NMR, Plenum, New York, with permission.)
the screening tensor, which has nine components, though only three
influence the observed screening. These, the diagonal components of
the tensor, are called crn, cr22 and cr33 and all three are required
if the nucleus is in a site with no symmetry. If the site is axially
symmetric, then only two values are needed to describe the screening
since an = a22. These components are then called a1 and a33 is denoted
a,,. This is called the screening anisotropy, and the differences between
the values of the components can be substantial.
Screening anisotropy of a nucleus can be observed by taking its
spectrum from a powdered solid where the solid particles all have
different orientations. Two types of spectrum are shown diagram-
matically in Fig. 2.3, and the values of the tensor components are
given by the discontinuities on the curves. The effect is another
source of line broadening in solids but also allows a full description
of the screening mechanism to be obtained. In liquid samples,
the isotropic rotation of the molecules produces an average a where
a = (an + a22 + a33)/3 and the lines become narrow. The high-
resolution spectrum that is thus obtained is generally more useful to
the chemist, but it should not be forgotten that some fundamental
information is lost in the process.
www.pdfgrip.com
Effects due to the molecule 21
(a) (b)
www.pdfgrip.com
22 The magnetic field at the nucleus
www.pdfgrip.com
Effects due to the molecule 23
for instance, can support a large electronic ring current around the
conjugated ir-bond system when the plane of the ring is transverse to
the field axis but very little when the ring lies parallel to the field axis.
This results in large average descreening of benzene protons since the
average secondary magnetic field, which must oppose the applied field
within the current loop, acts to increase the field outside the loop in
the region of the benzene protons (Fig. 2.6).
(a) (b)
Figure 2.6 Ring current descreening in benzene, (a) The plane of the benzene
ring is at right angles to the applied magnetic field B0. A current is induced
in the delocalized Tr-orbitals generating an opposing magnetic field in the
centre of the ring and a reinforcing magnetic field outside the ring. The result
is that protons above the ring move to lower frequency, while those outside
the ring, in the plane of the ring, move to higher frequency, (b) When the
applied magnetic field lies in the plane of the benzene ring, the area of the
ring current loops due to the electrons is much smaller.
www.pdfgrip.com
24 The magnetic field at the nucleus
The closer the protons are to the bond generating the electric field,
then the more they are descreened. This also shows up as a fall-off in
descreening along an alkyl chain for the protons further away from
the substituent, and, for instance, in 1-chloropropane (n-propyl chlo-
ride) the comparable descreening figures are a-CH2 3.24 ppm, (3-CH2
1.58 ppm and CH3 0.83 ppm.
Because the bonds in molecules are not rigid fixed entities but have
dimensions determined by vibrational phenomena, the substitution of
an atom by one of its isotopes of different mass alters the vibrational
energies in the bonds to that atom and so alters the electron distrib-
ution about it. This necessarily implies changes in nuclear screening
following the substitution, both near the substitution site and at some
distance from it. Such changes are small but measurable in many cases,
and provide useful spectroscopic data. It is necessary to distinguish
between primary isotope effects and secondary isotope effects. The
former are the effects experienced by the isotopically substituted
nucleus itself. For instance, for 14N/15N substitutions, the change in 14N
screening between 14NH3 and CH314NO2 is not exactly the same as that
for 15N between 15NH3 and CH315NO2, given identical conditions.
Clearly, such changes are not easy to measure and for many elements
are within experimental error, so that primary effects will not be
considered further. The secondary effects are those observed on the
nuclei in the rest of the molecule; in the above example the proton
screening changes, say, between 14NH3 and 15NH3. The trends observed
are as follows.
www.pdfgrip.com
Isotope effects 25
[15N1802r
[l5N16o18or
I I I I I I I
0.2 0.0 -0.2 -0.4 -0.6
8/PPM
Figure 2.7 The 15N NMR spectrum of the nitrite ion, [NO2]~, in water. The
ion was enriched to 95% in 15N and 1577% 16
in 1815O. 16The18 three resonances
15 18
arise
from ions of isotopic composition N O2, N 16 O O and N O2 in the
concentration ratios 6:33 :61. Replacement of O by 18O causes a low
frequency shift of 0.138 ppm. (From Van Etten and Risley (1981) /. Am.
Chem. Soc., 103, 5634; copyright (1981) American Chemical Society, reprinted
with permission.)
www.pdfgrip.com
26 The magnetic field at the nucleus
structure. It is, for instance, now possible to count the number of water
molecules in aqua complexes such as [A1(H2O)]3+. Analytical methods
never give whole numbers, and while the hexaaqua cation is known
to exist in solids from X-ray structural analysis, it is possible to argue
that the hydration number may not be so definite in solution. If
A1(H2O)6(C1O4)3 is dissolved in acetone, it is possible to observe the
water proton signal and note that this is highly descreened relative
to free water. This is a consequence of the strong electric field of
the cation. If a proportion of the water is replaced by D2O, then the
complex will contain H2O, HOD and D2O. The signal of the HOD
protons shows a strong isotope effect, though this is abnormal as they
are less screened even than the H2O. This probably arises because the
deuterium substitution shortens the Al-O bond in that molecule and
so increases the electric field effect.
More importantly, the two proton resonances (HOD and HOH)
show fine structure as in Fig. 2.8. This arises because of long-distance
Figure
2
2.8 (a) The 400 MHz 1H NMR spectrum of the water complexed to the cation [A1(H2O)]3+ in
( H6)acetone (deuterioacetone or acetone-d6) taken at -30°C. The complex had been partially deuteriated
so as to contain 35% 2H. The resonance at 10.23 ppm is due to all the HOD molecules in the complex
and that at 10.17 ppm is due to all the HOH molecules. The fine structure arises because different mole-
cules have different total numbers of 2H, each giving a smaller isotope effect due to the more distant
substitution. The stick diagram gives the calculated intensities obtained from the deuterium content and
assumes completely random distribution throughout the sample. (After Akitt et al. (1986) /. Chem. Soc.
Chem. Commun., 1047, with permission.) (b) The 95.7 MHz 59Co NMR spectrum of [Co(NH3)6]3+ in various 3+
mixtures of D2O/H
3+
2O to generate the complete range of nineteen deuterated compounds from [Co(NH3)6]
to [Co(ND3)6] . The introduction of each deuterium moves the signal approximately 6 ppm to low
frequency, (i) 15% D2O/85% H2O, (ii) 50% D2O/50% H2O, (iii) 85% D2O/15% H2O. The number besides
each signal refers to the number of hydrogen atoms that have been replaced by deuterium. (Remeasured,
but based on Russell and Bryant (1983) Anal Chim. Ada, 151, 227, copyright (1994), with permission
from Elsevier Science.)
www.pdfgrip.com
Effects due to unpaired electrons 27
isotope effects between a given HOD and the other ten replaceable
sites in the complex. A 13-line pattern is theoretically possible, though
some lines are too weak to detect, and one has to calculate an inten-
sity distribution from the known level of deuteration.
The changes can be even more dramatic when a nucleus with a
larger chemical shift range is observed. This is shown in Fig. 2.8(b)
for 59Co in [Co(NH3)6]3+, where replacement of hydrogen by deuterium
produces a shift of about 6 ppm for each replacement.
The electron (spin = 1/2) has a very large magnetic moment and if,
for instance, paramagnetic transition-metal ions are present in the
molecule, large effects are observed. The NMR signal of the nuclei
present may be undetectable, but under certain circumstances, when
the lifetime of the individual electron in each spin state is short, so
that its through-space effect averages to near zero, NMR spectra can
be observed. The screening constants measured in such systems,
however, cover a very much larger range than is normal for the
nucleus, and this arises because the electronic spins can be apparently
delocalized throughout a molecule and appear at, or contact, nuclei.
The large resonance displacements that result are known as contact
shifts and the ligands in certain transition metal-ion complexes
exhibit proton contact shifts indicating several hundred ppm changes
in a. In addition, if the magnetic moment of the ion is anisotropic,
one gets a through-space contribution to the contact interaction similar
to the neighbour anisotropy effect, and this is called a pseudo-contact
shift.
The NMR signals are also broadened by the presence of the
unpaired electron(s). The broadening is proportional to y2/r6, where y
is the gyromagnetic ratio of the nucleus being observed and r is the
distance between the unpaired electron(s) and the observed nucleus.
www.pdfgrip.com
28 The magnetic field at the nucleus
I I I I I I
140 120 100 80 60 40 20 0
1
H/ppm
1
Figure 2.9 The 100 MHz H NMR spectrum of [Co(4,6-Me2-phenanthro-
line)3]2+, (2.3), in CD3OD at -20°C. Note that the compound consists of two
isomers, mer and fac. (Reproduced by permission of the American Chemical
Society from La Mar and Van Hecke (1970) Inorg. Chem., 9, 1546.)
for the H2 and H9 signals which are moved by some HOppm. Note
that despite the presence of the unpaired electron, the signals are rela-
tively sharp.
However, in other cases, the 1H NMR signals may be so broad
that separate signals cannot be resolved. Even then all is not lost.
As the broadening depends on y2, the solution to the problem is to
observe a low y nucleus. 7(2H)2 is only 0.024 that of :H resulting in
much sharper signals, and the result is that deuterated samples of Cr3+
and Cu2+ compounds which fail to give usable !H NMR spectra can
give useful information in their 2H NMR spectra. For example, in
[Cr3(M,-OH)2(ji-02CCR3)4(02CCR3)2(bipy)?][C104], R - H or D, the
resolution of the acetate signals is greatly improved when R = D and
the 2H NMR spectrum is recorded compared with the 1H NMR spec-
trum, when R = H (Fig. 2.10), where all the acetate signals are observ-
able giving five peaks in the ratio 2:1:1:1:1.
www.pdfgrip.com
Effects due to unpaired electrons 29
3D2HCN
CDHoCN
40 30 20 10
2
H/ppm
50 0 -50
1
(a) H/ppm (b)
Figure 2.10 The 360 MHz 1H, (a), and 76.75 MHz 2H, (b), NMR spectra of [Cr3(fji-OH)2(^-O2CCR3)4
(O2CCR3)2(bipy)2][ClO4], R - H or D, in CD3CN, (a), or CH3CN, (b). In spectrum (b), only the signals
due to the deuteriated acetate ligands and the residual deuterium in the solvent are observed. Note the
better resolution observed for the acetate signals at 16 and 42ppm in the 2H NMR spectrum when
compared with the !H NMR spectrum. The numbered resonances in (a) arise from the bipyridyl ligand.
(Reproduced by permission of the American Chemical Society from Harton et al (1997) Inorg them, 36,
4875.)
www.pdfgrip.com
30 The magnetic field at the nucleus
1
20 10 0 -10 -20 -30 H/ppm
Figure 2.11 (a) 80MHz *H NMR spectrum of [(Ti5-C5H5)3UC(O)CH2CH2
CH2CH3] in C6D6. (b) 1H NMR spectrum of a mixture of [(T]5-C5H5)3UC
(O)CH2CH2CH2CH3] and [(ti5-C5H5)3UCH2CH2CH2CH3] in C6D6, the latter
giving the primed signals. (Reproduced from Paolucci et al (1984) /.
Organomet. Chem., 272, 363, copyright (1984), with permission from Elsevier
Science.)
www.pdfgrip.com
Effects due to unpaired electrons 31
<al
(b)
(c)
5 4 31 , 2 1 0
H/ppm
Figure 2.12 Three 400 MHz 1H NMR spectra of 1-pentanol in CDC13. (a)
With added [Eu(fod)3]. (b) No added shift reagent, (c) With added [Pr(fod)3].
Note that fod is 6,6,7,7,8,8,8-heptfluoro-2,2-dimethyl-3,5-octanedionate (2.4).
Signals due to the fod ligand are also observed.
www.pdfgrip.com
32 The magnetic field at the nucleus
(C)
(b)
(a)
Figure 2.13 The 2H NMR signals of the CHD deuterons in various samples
of the optically active hexanol, C5HUCHDOH, made using an optically active
reducing agent and obtained in the presence of a chiral shift reagent, in which
the ligands are hexafluoropropyl camphor ate (see formula). The lower trace
shows the racemic mixture and the two upper traces are from materials made
from reducing agents of opposite chirality.
whence
V l _ V 2= ^(a2-ai) (2.5)
www.pdfgrip.com
The chemical shift 33
^ = .,-., (17)
V
l
www.pdfgrip.com
34 The magnetic field at the nucleus
TMS
8
- T«ilf394 x W - W S f f m M
Or, more easily remembered, it is the frequency difference in hertz
divided by the frequency of the reference in megahertz. We see also
that the frequency separations will be different in spectrometers oper-
ating at different frequencies. Thus in a 200 MHz spectrometer, the
frequency separation above will be 1450 Hz.
The conventions adopted for other nuclei are less firm. The shifts
are usually large, so that it is not quite so important to be able
to compare different workers' results with high accuracy, and the
standard substance is often chosen according to the dictates of
www.pdfgrip.com
Sample preparation, standardization, solvent and temperature effects 35
H bonded
e.g. CHC13
'Hppm HF; C 6 H 6 C 2 H 4 (CH3)2O (CH3)2CO CH4 TMS(O) MH
. 1 . 1 _J ^ .
to-20 7.2 5.5 3.2 2.1 0.2 Oto-50
MeOH
MnO; CrO^~ Me2CO Ni(CO)4 SO^- H2O j Me2O
17
0 ppm ^ 1 1 1 1 1 T~^
1230 835 569 362 167 0 -53
Heteropolyanion oxygens span whole range
Figure 2.15 Various chemical shift scales and chemical shifts of some compounds.
www.pdfgrip.com
36 The magnetic field at the nucleus
convenience. The standard is assigned 0 ppm and the 8 scales are then
as in Fig. 2.15. An aqueous salt solution is often used as standard for
groups 1, 2, 3, and 17: 7Li+ (aq) for lithium, [27A1(H2O)6]3+ for
aluminium and 35C1~ (aq) for chlorine, for instance. Some much used
references are: (CH3O)3B and (CH3CH2)2O -»BF3 for n B spec-
troscopy; nitromethane, CH3NO2, or nitrate ion, [NO3]~, for 14N and
15
N spectroscopy; H2O for 17O spectroscopy; 85% orthophosphoric
acid, H3PO4, for 31P spectroscopy; and the refrigerant CFC13 is
commonly used as standard for 19F work. This 19F chemical shift scale
is often called the 4> scale. Other standards used for 19F spectroscopy
are hexafluorobenzene, C6F6, or trifluoroacetic acid.
These and some other chemical shift scales are illustrated in Fig.
2.15, which shows the standards used (0 ppm) and the chemical shifts
of some compounds, chosen to cover the full range of shifts for a given
element rather than to pick out any particular trends with composi-
tion. For many of the heavier elements, the shift is very medium sensi-
tive and the spot value given is purely illustrative. For the 19F scale,
the shift marked B is that of the bare nucleus (189 ppm). Note also
the very large chemical shift range for the transition-metal nucleus
59
Co.
www.pdfgrip.com
Sample preparation, standardization, solvent and temperature effects 37
setting the NMR tube in the spinner at the same depth, using a depth
gauge. Provided enough solvent is used so that the top of the solution
is well above the coil, there will be only minor changes to the field
homogeneity, and this is the normal procedure for samples with strong
signals. Where sensitivity is low and the quantity of compound is
limiting, it is advisable to have as much compound as possible inside
the detector coils. This is done by adjusting the position of the bottom
of the tube and the top of the solution to match the bottom and top
of the detector coil, but considerable time will be required to reshim
the magnet to obtain best homogeneity.
2.5.1 Standardization
We have shown that chemical shifts are invariably measured relative
to a standard of some sort. There are five ways of standardizing a
resonance, which are now given.
www.pdfgrip.com
38 The magnetic field at the nucleus
www.pdfgrip.com
Sample preparation, standardization, solvent and temperature effects 39
8
int = 8obs - -y (Xref ~ Xsample) (211)
8
int = 8obs + -y (Xref ~ Xsample) (2'12)
where 8int is the corrected chemical shift, S0ts and Sotjsare the measured
chemical shifts, xref and Xsample are tlie magnetic susceptibilities of the
reference capillary and the sample, respectively. Most of the results
quoted in the literature using an external reference have not been cor-
rected for susceptibility effects, e.g. for 31P, an external reference of
85% H3PO4 is normally used, and the result is a discrepancy of up to
1 ppm between the older work recorded using a permanent or elec-
tromagnet and the more recent work using a superconducting solenoid.
Because the capillary holding the standard distorts the magnetic field
around it, the field homogeneity in the annular outer part of the sample
www.pdfgrip.com
40 The magnetic field at the nucleus
www.pdfgrip.com
Sample preparation, standardization, solvent and temperature effects 41
solvent shift. For this reason, a complex solute may have quite different
spectra in, for instance, chloroform and benzene, and a change from
one solvent to the other may remove some degeneracy or avoid the
overlapping of signals that would otherwise be difficult to disentangle.
Figure 2.16 shows the quite large changes that can occur. This tech-
nique is often known by the acronym ASIS, or Assisted Solvent-
Induced Shifts.
Chloroform as a solute suffers considerable solvent shifts. The pure
liquid is self-associated by hydrogen bonding, but upon progressive
dilution in an inert solvent the proportion of hydrogen-bonded mol-
ecules is reduced and its resonance is shifted 0.29 ppm to low
frequency. The shifts noted in Table 2.1 are in excess of this and we
must consider the existence of several other types of interaction. Thus
specific interactions with the Lewis bases, dimethylsulfoxide and
pyridine, result in high-frequency shifts. In the case of pyridine, this
implies a preference for an edge-on approach to the aromatic ring and
(d)-
(c)-
(b)-
(a) =
2.4 2.2 2.0 1.8 1.6 1.4 1.0 0.8 0.6 0.4
1
H/ppm
Figure 2.16 400 MHz 1H NMR spectra of camphor, (a) Using CDC13 as solvent, (b) As (a), but with
increased gain, (c) Using C6D6 as solvent, (d) As (c), but with increased gain. Note that in CDC13, the
signals due to H5a and H6a overlap, while they are separated in C6D6. However, in C6D6, the signals due
to H3a, H4, and H5b now overlap. The positions of the signals are markedly solvent dependent. The fine
structure is due to spin-spin coupling, to be discussed in the next chapter.
www.pdfgrip.com
42 The magnetic field at the nucleus
www.pdfgrip.com
Questions 43
2.6 QUESTIONS
www.pdfgrip.com
44 The magnetic field at the nucleus
given in Fig. 2.15. Use the data to estimate <rd for 59Co. Assume
that AE1'1 is the only significant term contributing to the ap
variations.
2.5. In annulene,
the inner protons have a chemical shift of 8 -3.0 while the outer
ones come at 8 9.3. Explain why this happens when the XH chem-
ical shift of ethene is 8 5.5.
2.6. Figure. 2.17 shows the 195Pt NMR spectrum of [PtCl6]2-. Explain
why the signal is a multiplet and account for the number of signals
and their intensities.
2.7. Many old reports involving the use of NMR spectrometers with
electromagnets report the 31P chemical shift of PPh3 as being
between 8 -6 and -7. More recent reports using NMR spec-
trometers with superconducting magnets report the 31P chemical
shift of PPh3 as being between 8 -5 and -6. Explain why the
apparent chemical shift has changed.
2.8. E for 125Te in the reference compound Me2Te has been reported
as being 31.549 802 MHz. If TMS in a sample in CDC13 has been
measured as being at 400.134 394 MHz, what frequency should
be used to reference a 125Te signal from this solution with Me2Te
I at zero? The 125Te NMR signal of a sample of Ph2Te was deter-
100 OHz mined as being at 126.365 452 MHz. What is the chemical shift
in ppm relative to Me2Te?
Figure
195
2.17 The 85.6 MHz 2.9. There are two common references for 195Pt, namely E = 21.4 MHz
Pt NMR spectrum of and [PtCl6]2~ which has E = 21.496 700 MHz. The compound
2M Na2[PtCl6] in D2O. [Pt(CN)4]2- has E = 21.395 634 MHz. Calculate the chemical shifts
(Reproduced by permis-
sion of The Royal Society of [PtCl6]2- and [Pt(CN)4]2- using first E = 21.4 MHz and then
of Chemistry from Sadler [PtCl6]2~ as the reference. Explain why when using equation (2.6),
et al (1980) /. Chem. Soc., the chemical shift separation of [PtCl6]2~ and [Pt(CN)4]2~ in ppm
Chem. Commun., 1175.) depends on the chosen reference.
2.10. Using equations (2.11) and (2.12), work out an expression which
allows the magnetic susceptibility differences between two liquids
to be obtained from measurements of the chemical shifts of reso-
nances in the two liquids in both a superconducting magnet and
an electromagnet.
www.pdfgrip.com
Intei nuclear spin-spin
coupling s
3.1 THE MUTUAL EFFECTS OF NUCLEAR MAGNETS ON
RESONANCE POSITIONS
www.pdfgrip.com
46 Internuclear spin-spin coupling
www.pdfgrip.com
The appearance of multiplets arising from spin-spin coupling 47
15-
N
I 10-
if
X
CO 5-
0-
0 45 90 135 180
0
Torsion Angle/
Figure 3.1 Karplus curves, using the equation modified by Altona, relating the dihedral angle 4> in a
HC-CH fragment and the vicinal proton-proton coupling constant. The inset shows a view along the
carbon-carbon bond. Two curves are shown relating to differently substituted molecules, CH3CH3, dashed
line, (a) and CH3CH2F, solid line, (b). The horizontal lines, (c) and (d), show the average value obtained
when a group can rotate freely, giving rise to an averaged /vic.
show quite clearly what the conformation of the molecule studied has
to be. Thus, in Fig. 3.2, part of the proton spectrum of menthol is
given in which various doublet splittings due to spin-spin coupling are
evident. The resonance of H1 is split into a doublet of triplets with a
small / of 4.2 Hz and two large ones of 10.5 Hz. H1 is easily identifi-
able by its chemical shift, being next to the electronegative OH group,
and it is the large couplings J(Hl-H2ax) and /(H^H6) that are due to
these axial-axial pairs with a dihedral angle of near 180° and large
predicted /. The other lesser interaction is /(ff-H2^), H1 and H2 being
an axial-equatorial pair with dihedral angle of near 60° and so small
predicted /.
This vicinal 3J dependence on dihedral or torsion angle seems to be
quite general and Karplus-type curves have been established for the
coupling paths ^C-C-C^H, 31P-C-C-31P and 13OC-C-31P and are
also likely for coupling between 77Se or 125Te and 1H or 13C or for the
more exotic systems such as 3/(199Hg-13C) in alkylmercury compounds.
www.pdfgrip.com
48 Internuclear spin-spin coupling
Figure 3.2 The 400 MHz !H NMR spectrum of menthol whose structure is shown in the inset. The numbers
represent hydrogen atoms attached to the ring. The resonances are split by spin coupling with adjacent
hydrogen atoms. In particular, H1, at 8 3.4, is split into a doublet, 4.2 Hz, of triplets, 10.5 Hz, by coupling
with the equatorial H6 and the two axial H2 and H6 nuclei.
www.pdfgrip.com
The appearance of multiplets arising from spin-spin coupling 49
www.pdfgrip.com
50 Internuclear spin-spin coupling
Figure 3.3 A simulated signal demonstrating the splitting due to two / = 1/2
nuclei. When Sra = 0 there is no perturbation of the coupled resonance so
that the centre line corresponds to the chemical shift position. This holds for
all multiplets with an odd number of lines. The spacing / between the middle
of the lines corresponds to a change in Sra of unity.
2m = + -i- + -L = +1 (3.1)
or
f +2^ r - 42 - = o
Sm = J
(3.2)
1-1 + 1 = °
^-i-i.-i
or
(3.3)
www.pdfgrip.com
The appearance of multiplets arising from spin-spin coupling 51
triplet. The splitting of the CH2X resonance caused by the CH3 group
can be found from Fig. 3.5, and is a 1 : 3 : 3 :1 quartet. A typical ethyl
group spectrum is shown in Fig. 3.6.
We have done enough now to formulate a simple rule for splitting
due to groups of / = 1/2 nuclei. Thus the number of lines due to
coupling to n equivalent / = 1/2 nuclei is n + 1. The intensities of the
lines are given by the binomial coefficients of (a + 1)" or by Pascal's
triangle, which can be built up as required. This is shown in Fig. 3.7.
A new line of the triangle is started by writing a 1 under and to the
left of the 1 in the previous line and then continued by adding adja-
cent figures from the old line in pairs and writing down the sum as
shown. The multiplicity enables us to count the number of I = 1/2
nuclei in a group and the intensity rule enables us to check our assign-
ment in complex cases where doubt may exist, since the outer compo-
nents of resonances coupled to large groups of nuclei, e.g. the CH of
(CH3)2CH may be too weak to observe in a given spectrum.
Coupling to nuclei with / > 1/2 leads to different relative intensities
and multiplicities. In the case of a single nucleus the total number of
spin states is equal to 21 + 1 and this equals the multiplicity. If / = 1/2
www.pdfgrip.com
52 Internuclear spin-spin coupling
we get two lines, 1=1 gives three lines, / = 3/2 gives four lines, and
so on. The spin populations of each state are virtually equal and so
the lines are all of equal intensity and of equal spacing (Fig. 3.8).
Figure 3.5 The splitting due to three / = 1/2 nuclei. There is no line that
corresponds to Sra = 0. The multiple! is, however, arranged symmetrically
about the Sra = 0 position, so that the centre of the multiplet corresponds to
the chemical shift. This rule holds for all multiplets with an even number of
lines.
JL J_
—r — IMS
I 3.5 3A 1.7 I
www.pdfgrip.com
The appearance of multiplets arising from spin-spin coupling 53
AT
0 1 Singlet
1 1 1 Doublet
+
2 1 2 1 Triplet
+ +
II II
3 1 3 3 1 Quartet
+ + +
4 1 4 6 4 1 Quintet
Figure 3.7 Pascal's triangle can be used to estimate the intensities of the lines
resulting from coupling to different numbers, N, of equivalent / = 1/2 nuclei.
The numbers in each line are obtained by adding adjacent pairs of numbers
in the line above.
/e.g.
\ 3lPMe3
2
J(31P1H) = 30Hz
1 [14NH4]+
1
J(14N1H) = 52.5Hz
f [UBH4]-
1
J(11BH) = 80.5Hz
3 [10BH4]-
1
J(10BH) = 27.2Hz
www.pdfgrip.com
54 Internuclear spin-spin coupling
www.pdfgrip.com
The appearance of multiplets arising from spin-spin coupling 55
Thus the rule given for 7 = 1/2 nuclei can be generalized to include
groups of nuclei of any given /, the number of lines observed for
coupling to n equivalent nuclei of spin / being 2nl + 1. Obtaining the
relative intensities is, however, tedious and it is better to use a 'Pascal's'
triangle type of construction as shown in Fig. 3.10 for when 1=1 and
Fig. 3.11 for when 7 - 3/2.
In the case of / = 1, the numbers in the triangle are placed immedi-
ately under the previous ones, since there is always a line in the centre
of the multiplet. For / = 3/2 (and all half-integral spins), the numbers
are staggered, since only if n is even is there a line at the centre of the
multiplet. The triangles are constructed by moving a box, which can
enclose 27+1 numbers of a line, along the line, enclosing first the 1,
then two numbers, then three, then four, and so on. The numbers
enclosed by the box are added to give a number for the next line. Thus
for 7 = 3/2 the box can enclose four numbers, and if we sweep through
the line for n = 1, we enclose first a 1, then 1 + 1 = 2, then 1 + 1 + 1
= 3, then 1 + 1 + 1 + 1 = 4 and reducing again to 3, 2, 1. The construc-
tion given above for 7 = 1/2 nuclei is the first example of this series.
N 7=1
0 1
1 1 1 1
2 1 2 3 2 1
3 1 3 6 7 6 3 1
1
4 1 4 10 16 19 16 10 4 1
Figure 3.10 A version of 'Pascal's triangle' to determine the intensities of
lines resulting from coupling to different numbers, N, of equivalent 7 = 1
nuclei.
N 7=3/2
0 1
1 1 1 1 1
2 1 2 3 4 3 2 1
3 1 3 6 10 12 12 10 6 3 1
i
4 1 4 10 20 31 40 44 40 31 20 10 4 1
Figure 3.11 A version of 'Pascal's triangle' to determine the intensities of
lines resulting from coupling to different numbers, N, of equivalent / = 3/2
nuclei.
www.pdfgrip.com
56 Internuclear spin-spin coupling
Coupling to HX then
gives a doublet of
doublets
Figure 3.12 Splitting of the H¥ protons of Z2CHA-C(HM)2-CHXY2 due to coupling to HA and Hx. (a)
/(A-M) = /(X-M), the centre lines overlap and the multiplet is a 1: 2 :1 triplet just as if HM were coupled
to a CH2 group, (b) /(A-M) * /(M-X) and we get a doublet of doublets from which both /(A-M) and
/(M-X) can be measured.
www.pdfgrip.com
The appearance of multiplets arising from spin-spin coupling 57
the nuclei of the BH4 ligand only, with /(9Be-H) - 10.2 Hz and /(9Be-
n
B) = 3.6 Hz. The first coupling causes splitting to a 1 : 4 : 6 : 4 : 1
quintet, each line of which is further split into a 1:1 :1 :1 quartet by
the nB. So 20 lines are expected, but because the ratio of the coupling
constants is almost integral (it is 2.8) some lines overlap and only 16
are observed. The overlap is not exact but the close pairs of lines are
not resolvable because of the appreciable natural linewidth of the 9Be
resonance.
3.2.1 Rules for the analysis of any first-order multiplet arising from
7 = 1 / 2 coupling
The analysis of any well resolved first-order multiplet due to spin 1/2
nuclei is actually a trivial exercise. There are some very simple rules
which make the analysis straightforward.
1. Check that the multiplet is centro-symmetric. If it is not, then it
consists of either overlapping multiplets or is second order, see
section 3.7.
2. Assign an intensity to each line in the multiplet, starting with 1 for
the outermost lines and using only integers. The distribution of the
intensities must be centro-symmetric.
3. Add up the assigned intensities. The sum must be 2", where n is
the number of / = 1/2 nuclei coupling. If the sum does not come
www.pdfgrip.com
58 Internuclear spin-spin coupling
www.pdfgrip.com
The appearance of multiplets arising from spin-spin coupling 59
www.pdfgrip.com
60 Internuclear spin-spin coupling
The intensities are put on each line and they sum to 16 = 24. Hence
there are four / = 1/2 nuclei coupling. The separation of the first pair
of lines is a coupling constant, and their intensities are 1 :2. Hence,
from 'Pascal's triangle', this is part of a 1:2:1 triplet. The separation
between this pair of lines must equal that between the central one and
the third line of intensity one, so the first, second, and fourth lines are
selected as making up the triplet. The triplet is drawn above the multi-
plet. We now move to the next unassigned line, the third one and
repeat the same triplet, using the same coupling constant. We keep
repeating this until all the lines are accounted for. A 1:1 :1 :1 quartet
has been generated. This rapidly reduces to a doublet, and then a
singlet and the pattern is analysed.
3.2.2 Rules for the analysis of any first order multiple! arising from
both 7 = 1 / 2 and / > 1/2 coupling
The rules for determining the sum of intensities of the lines need modi-
fying for nuclei with 7 > 1/2.
i
Sum of intensities = JJ (21 j + 1)"'
;=o
where there are n^ nuclei with spin 7;. The appropriate intensity
triangle for the nuclear spin has to be used. We return to the case of
[(^5-C5H5)BeBH4] to illustrate this (Fig. 3.14). The intensity ratio
of the lines is 1:1 :1 : 5 : 4 : 4 :10 : 6 : 6 :10 : 4 : 4 : 5 :1 :1:1. (Note
that the intensities are not obvious from the spectrum in this case
due to the overlap of the lines.) The sum of the intensities is 64. The
coupling is to four protons and one nB. The sum of intensities should
therefore be (2 x 3/2 + 1)(2 x M + I)4 = 4 x 24 = 64.
The separation of the outer pair of lines is a coupling constant. This
could be to either n B or to 1H, but coupling to the four protons gives
a 1 : 4 : 6 : 4 : 1 quintet, and the multiplet starts 1:1:1. Hence the
smallest coupling constant is to nB, giving a 1:1 :1:1 quartet. This
1:1:1:1 pattern is drawn together by drawing a vertical line above
the middle of it and the intensity of the outermost line, 1, is attached.
The 1:1:1:1 intensity is taken away from the first four lines to leave
residual intensities 0 : 0 : 0 : 4 : 4 : 4 :10 : 6 : 6 :10 : 4 : 4 : 5 :1 :1:1. The
same pattern is subtracted, but the intensity starts 4, so the intensities
to be subtracted are 4 : 4 : 4 : 4 , and 4 is written beside the central
vertical line. These intensities are then subtracted to leave residual
intensities 0 : 0 : 0 : 0 : 0 : 0 : 6 : 6 : 6 :10 : 4 : 4 : 5 :1:1:1. The same
pattern is subtracted again, but the intensity starts 6, so the intensi-
ties to be subtracted are 6 : 6 : 6 : 6 , and 6 is written beside the central
vertical line. These intensities are then subtracted to leave residual
intensities 0 : 0 : 0 : 0 : 0 : 0 : 0 : 0 : 0 : 4 : 4 : 4 : 5 : 1 : 1 : 1 . The procedure
is continued reducing the multiplet to a 1:4:6:4:1 quintet, consistent
with four protons coupling.
www.pdfgrip.com
Spin-spin coupling satellites 61
A glance at the table of nuclear properties (Table 1.1) will show that
certain elements have as principal isotope a magnetically non-active
species, but that they have also a more or less small proportion of
magnetically active species, some with 7 = 1/2. Examples are platinum
with 33.8% of the active nucleus 195Pt and carbon with just 1.1% of
the active 13C. The spectra of other nuclei in compounds of these
elements will thus arise from differentiable molecular species: those
with the non-active isotope, which will give rise to intense patterns
depending upon the other nuclei present, and those with the active
isotope, which will give a weaker sub-spectrum in which spin-spin
coupling will be seen arising from interaction with this less abundant
isotope. These weak doublets are centred approximately around the
corresponding lines in the spectrum of the main species and are thus
called satellites. They are usually particularly obvious in !H or 31P
spectra of platinum compounds, since their intensity is about a quarter
of that of the central line of the major species. We will give two exam-
ples of compounds containing one and two platinum atoms.
First, we will consider the proton spectrum of the mononuclear
square planar complex ion, [PtMe(PMe2Ph)3]+. The methyl groups
attached to platinum and to phosphorus are both coupled to 195Pt, and
their spectra, in the low frequency region, are shown in Fig. 3.16, with
the effect of the 31P atoms removed by a process of double irradia-
tion, which we will discuss in detail in Chapter 5. There are three
major methyl resonances which originate from the complexes with
magnetically inactive platinum, and each is flanked by two lines of
one quarter intensity, which are the doublets due to coupling to one
www.pdfgrip.com
62 Internuclear spin-spin coupling
(a)
40 20 0 -20 6(31P)
(b)
(c)
11 11
(d)
Figure 3.17 The 31P NMR spectrum of [Pt2(CO)2(PPh3)2(fi-MeO2CC>
CCO2Me)], (3.1), with the effect of the !H nuclei removed by a double
irradiation technique, (a) The experimental spectrum with signals due to
coupling to 195Pt, I = 1A, shown above with increased gain, (b) The simulated
spectrum for the molecules containing no 195Pt. (c) The simulated spectrum
for the molecules containing one 195Pt atom, (d) The simulated spectrum for
the molecules containing two 195Pt atoms. (Reproduced with permission from
Kose et al (1981) Inorg. Chem., 20, 4408, copyright (1981) American Chemical
Society.)
www.pdfgrip.com
Spin-spin coupling satellites 63
into doublets, to give eight lines in all. The remaining 11.4% of the
molecules contain two 195Pt atoms and give a weak spectrum. This sub-
spectrum is a second-order spectrum type, [AX]2, to be discussed in
section 3.5.4. Analysis of this sub-spectrum yields 1/(195Pt195Pt) =
786 Hz. The spectra become much more complex for bigger clusters
and are of considerable use in studying the structures of this inter-
esting class of compound.
Another nucleus that provides interesting examples of spin coupling
satellites is 183W, which has a natural abundance of 14.28%. The 19F
spectrum of the binuclear complex [W2O2F9]~ is shown in Fig. 3.18,
and consists principally of a doublet of intensity 8 and a nonet of inten-
sity 1. The nine fluorine atoms can thus be divided into an isochro-
nous set of eight atoms and one unique atom. The main spectrum
arises from those molecules in which both tungsten atoms are the
magnetically inactive tungsten isotopes but satellite lines are also
observed due to the molecules with one 183W atom in their structure.
Figure 3.18 The 56.4 MHz 19F spectrum of [W2O2F9]-, (3.2), in (MeO)2SO,
showing spin satellites due to 14.28% of 183W. (a) The multiple! due to183the
terminal fluorides, (b) Expansions of the central signals which are the W
satellites, (c) The multiplet of the bridging fluoride recorded at higher gain,
(d) The structure of the molecule showing the various coupling constants in
hertz. The outer lines of the nonet are lost in the baseline noise and the
student should confirm that the intensity ratios of the observed lines corre-
spond to those expected for the inner seven of the nonet rather than to those
expected for a septet. (After McFarlane et al. (1971) /. Chem. Soc. A, 948,
with permission.)
www.pdfgrip.com
64 Internuclear spin-spin coupling
In principle, lines should also exist due to those molecules with two
183
W atoms, but their proportion is low and their resonances are too
weak to observe. Each of the lines of the intense doublet has two 183W
satellites, each of which is further split into a 1 : 4 : 6 : 4 : 1 quintet.
This pattern must arise from coupling to four fluorine atoms. We can
therefore conclude that we have four of the eight isochronous fluo-
rine atoms associated with the 183W atom and therefore split into a
satellite doublet and then further coupled to the remaining four, which
are equally associated with the NMR inactive tungsten isotopes. This
provides considerable confirmatory evidence that the structure is
[OWF4»F»WF4O]~ with a fluorine atom bridging the tungsten atoms.
Finally we must consider the effect that the 1.1% of naturally occur-
ring 13C has on the proton spectra of organic compounds, which contain
principally the inactive 12C. Some of the molecules will however
I /
7.0 6.9 6.8 6.7 6.6
6(1H)
Figure 3.19 The 400 MHz !H NMR spectrum of the olefinic protons of dimethyl fumarate,
MeO2CCH=CHCO2Me in CDC13. In addition to a spectrum at normal gain,13 above there is a spectrum
with 13increased gain to show the 13C satellites. The molecules, MeO 12
2C CH= CHCO2Me show coupling.
The CH proton is split into a large doublet, 1J(13C1H) = 167.9 Hz, which is further split by the 12CH
proton with 3J(1H1H) = 15.8 Hz. The 12CH proton is split into a small doublet, 2J(13C1H) = 5.9 Hz, which
is further split by the 13CH proton with 3J(1H1H) = 15.8 Hz.
www.pdfgrip.com
The description of spin systems 65
contain one 13C atom, distributed randomly. In fact, the spin satellites
due to this minor isotopic component are not easy to observe among
the intense JH-12C resonances, but have proved to be of considerable
use. In simple compounds such as acetone, (CH3)2CO, for instance,
the proton resonance has two pairs of spin coupling satellites due to
molecules with one 13C atom in the methyl group (1J(1H-13C) = 126 Hz}
and to those with 13C in the carbonyl group [2J(1H-13C) = 5.9 Hz}.
Thus we can measure proton-carbon coupling constants, and with
double-resonance techniques we will see later that we can discover
correlations between the proton and carbon spectra of a molecule, see
Chapter 8.
The reduction of symmetry in a molecule caused by the presence
of 13C can also prove valuable. A simple example is to determine
whether dimethyl fumarate, MeO2CCH=CHCO2Me, has a cis- or
frYws-configuration at the double bond. The compound gives two
singlets in the ratio 1 :3 in the 1H NMR spectrum due to the olefinic
CH and the methyl protons. In the 12C compounds, the olefinic protons
are equivalent and give a singlet. However, the presence of one olefinic
13
C carbon atom removes the equivalence and the coupling between
the two olefinic protons can be measured as 15.8 Hz (Fig. 3.19). The
15.8 Hz coupling is in the range to be expected for a trans-stereo-
chemistry about the double bond in view of the electronegativity of
the substituents.
www.pdfgrip.com
66 Internuclear spin-spin coupling
www.pdfgrip.com
The description of spin systems 67
50 Hz
(b)
XX
(a).
Figure 3.21 A partial 250 MHz !H NMR spectrum of Ph3SiCH2CH2SiMe3 in CDC13, showing the CH2CH2
protons, (a) Experimental spectrum, x = impurity, (b) Calculated spectrum. (Reproduced with permission,
V.E. McGrath, PhD thesis, Sheffield, 1993.)
www.pdfgrip.com
68 Internuclear spin-spin coupling
www.pdfgrip.com
Second-order effects 69
The rules so far discussed apply to spectra of nuclei of the same species
where the separation between multiplets in hertz (i.e. the chemical
shift) is large compared to the value of the coupling constant between
them, or to coupling between nuclei of different elements or isotopes
where the differences in NMR frequency are invariably large.
Second-order spectra arise when the frequency separation between
multiplets due to different magnetically equivalent sets of nuclei is
similar in magnitude to the coupling constant between them; under
these circumstances, the effects due to spin coupling and chemical shift
have similar energy and become intermingled, leading to alterations
in relative line intensities and in line positions. Because it is the ratio
between / and the frequency separation that is important, chemical
shifts are always expressed in hertz (Hz) and not in parts per million
(ppm) when discussing second-order spectra. The hertz separation is
obtained by multiplying the chemical shift 8 by the spectrometer oper-
ating frequency and is written v08. The perturbation of the spectra
from the first-order appearance is then a function of the ratio //v08
and is different for spectrometers operating at different frequencies.
If a high enough frequency is used, many second-order spectra
approach their first-order limit in appearance and are then much more
easily interpreted. This is one of the advantages of high-field instru-
mentation. We will describe first the simplest possible system consisting
of two 7 = 1/2 nuclei.
www.pdfgrip.com
70 Internuclear spin-spin coupling
=00
J- =
v08
(h)
- - =2.5
v08
(9)
1
vV
08
(f) JUUUl
-i = 0.866
<•>
- = 0.4
(d)
-4 =0 - 2
v06
(0
=0
v—s -1
08
(b)
-: =0.05
v08
(a)
1 1 1 1 1 1 1 1 1
www.pdfgrip.com
Second-order effects 71
Table 3.2 Wave functions and energy levels for the AB second-order spin
system
1 1
4
2 4
Table 3.3 Transition energies relative to the centre of the multiplet and
relative intensities for the AB second-order spin system.
1 1/ l J
a 3-^1 >)2 + J2 V(v0§)2 -h /
2
1 1
2/+2 y)2 + J2
1 ~\-
b 4-*2
°' V(v08)2 -h / 2
tf + J2
1 ~\-
c 2->l 2
V(v08) - h /
2
1 1 i ! J
d 4^3 J ))2 + J2 ± |
h/2
-2 ~2^
www.pdfgrip.com
72 Internuclear spin-spin coupling
(a) (b)
Figure 3.24 (a) The energy level diagram for a system of two spins related to the resulting AB quartet.
Note that the vertical scale is misleading. The energy gaps represented by the arrows are in MHz, while
the energy gap between energy levels 2 and 3 is in Hz. (b) The spectrum was calculated for / = 10 Hz,
v08 - 25 Hz, i.e. //v08 = 0.4.
The energy levels are marked with the appropriate spin state and the
transitions, which in the first-order case can be regarded as arising
from transitions of the A or B nucleus, form opposite sides of the
figure. The three line separations are a-b = c-d = J and b - c =
V(v08)2 + J2 - \J\. The separation a - c or b - d, which in the first-
order case is the same as the separation between the doublet
centres, and is therefore the chemical shift in hertz, v08, is now simply
V(v08)2 + J2 and is larger than the true chemical shift. In other words,
although v08 is reduced to zero, the doublet centres never coincide
and are separated by / Hz. The outer lines, however, have intensity
zero at this point, while the inner lines are coincident, i.e. we predict
a singlet spectrum as is observed (cf. Fig. 3.23(h)). Thus arises our
rule, 'isochronous coupled protons resonate as a unit'.
We can calculate some simple rules for analysing an AB spectrum:
1. The spectrum contains two intervals equal to /, a - b and c - d.
2. The true AB chemical shift v08 is found as
(a - d)(b -c) ={ V(v08)2 +72 + / } { VO^+T 2 - / }
= (v0S)2 + P-J2
= (v08)2
so
v08 = V ( a - < * ) ( f r - c ) (3.4)
where a - d is the separation between the outermost lines and
b - c is the separation between the innermost pair of lines.
3. The assignment can be checked against the intensity ratios of the
larger and smaller lines. The intensity ratio, stronger/weaker, is
www.pdfgrip.com
Second-order effects 73
1+
V(v08)2 + /2
J
1- -
V(v08)2 + ^2
which gives
V(v08)2 + /2 + /
V(v0s)2 + /2 - y
for the ratio of the line separations (a - d):(b - c). Note that
changing the sign of / does not alter the pattern.
Figure 3.23 shows the form of the AB quartet for several values of
//v08. It is important to remember that if a multiplet shows signs
of being highly second order then both intensities and resonance posi-
tions are perturbed from their first-order values. A spacing corre-
sponding to /AB remains in AB-type spectra since only one coupling
interaction exists, but in more complex systems the spacings are combi-
nations of coupling constants. On the other hand, if the intensity
perturbation is only slight {Fig. 3.23(a)} then the line positions are not
detectably perturbed.
Examination of the spectra in Fig. 3.23 show several features. The
first visible onset of second-order behaviour is the inner lines becoming
more intense. The two doublets are said to 'lean' towards each other.
This feature can be very valuable in deciding which multiplets are
coupled but care is necessary as the heights of signals can be distorted
by insufficient digitisation, see section 5.6.3. There are two features
which are traps waiting for the unwary and have led to the misinter-
pretation of spectra.
1. When the separation of the two signals is / s , the spectrum has
the appearance of a 1 : 3 : 3 : 1 quartet as if it arises from coupling
to a methyl group (Fig. 3.23(e)).
2. When //v08 is large, the outer lines are weak and can be lost in the
noise leaving an apparent doublet.
Four examples of actual spectra that contain an AB multiplet are
given in Fig. 3.25. We have however reached a point in the history of
NMR where, in many cases, the analysis of second-order spectra can
be avoided by operating at a high enough frequency. In those cases
where this will not work, computer programs are available that permit
the full analysis of virtually any system.
www.pdfgrip.com
74 Internuclear spin-spin coupling
7.9 7.8 7.7 7.6 7.5 7.4 7.3 7.2 7.1 7.0 6.9
1
(a) H/ppm
www.pdfgrip.com
Second-order effects 75
I I I I
55 50 45 40 35 30 25 20 15 10 5
31
(c) P/ppm
40 30 20
31
(d) P/ppm
www.pdfgrip.com
76 Internuclear spin-spin coupling
"B=^
J
_ \a - d + f-h\
AB ~ ~
Figure 3.26 A calculated AB2 spectrum with VA = 5 Hz and v# = -10 Hz, and
/AB = 4 Hz. Dotted lines are included to mark the chemical shifts of A and B.
www.pdfgrip.com
Second-order effects 77
1020 1010 1000 990 980 220 210 200 190 180
Hz Hz
Figure 3.27 Simulated ABX spectrum using 8(A) = 190 Hz, 8(B) = 210 Hz, 5(X) = 1000 Hz, /(AB)
10 Hz, /(AX) = 4 Hz and /(BX) = 13 Hz.
www.pdfgrip.com
78 Internuclear spin-spin coupling
www.pdfgrip.com
Second-order effects 79
1020 1010 1000 990 980 220 210 200 190 180
Hz Hz
Figure 3.28 Simulated ABX spectra using 8(A) - 197.72 Hz, 8(B) - 202.24 Hz, 8(X) - 1000 Hz, /(AB)
10 Hz, /(AX) = 11.4 Hz and /(BX) - -28.5 Hz.
www.pdfgrip.com
80 Internuclear spin-spin coupling
\J
1020 1010 1000 990 980 220 210 200 190 180
Hz Hz
Figure 3.29 Simulated ABX spectra using 8(A) - 190 Hz, 8(B) = 210 Hz, 8(X) - 1000 Hz, /(AB) - 10 Hz,
/(AX) - 4 Hz and /(BX) - -13 Hz.
(3.11)
Figure 3.30 The 100.62 MHz 13C NMR spectrum of the 13CH2 group of
[(Ti5-C5Me5)2Rh2(fjL-CH2)2{fju-CH2CH(CH2CH-CH2)CH2}], (3.11). The spec-
trum is ABX with the two 103Rh nuclei giving the AB part of the spectrum,
not shown, and the 13CH2 giving the X part. This spectrum has 1/(103Rh103Rh)
- 13.5 Hz, y^Rh^C) - 30.0 Hz, 2/(103Rh13C) - 0.0 Hz, and A8(103Rh) -
0.36 ppm or v08 = 4.5 Hz. (Reproduced with permission from Mann et al
(1985) /. Chem. Soc., Dalton Trans., 1555.)
www.pdfgrip.com
Second-order effects 81
www.pdfgrip.com
82 Internuclear spin-spin coupling
L = JAW-JAW
M= J
A1 A2 ~ J X1 X2
N = / A i x i + /AiX2
Each multiplet consists of an intense doublet, separation N, and two
AB patterns, apparent coupling constants K and M. The inner lines
of each AB pattern are separated by V^2 + L2 - K and VM2 + L2 -
M respectively.
TV can be easily determined, 34.6 Hz. K and M are 31.6 and 41.2 Hz.
The corresponding values of A/X2 + L2 - K and VM2 + L2 - M are 14.2
and 11.6 Hz. Substituting values of K and M yields L = 33.2 Hz. The
coupling constants are then obtained from combinations of K, L, M,
and N as follows.
7 K + M on „ TT
<W = —2— = 39.4 Hz
K
/Xlx2 = ~M = 4.8 Hz
A—I I iV
33
/-»/-% f\ f-w-
/ A i x i = —2— = -9 Hz
'A'X>=^^- = 0 - 7 Hz
www.pdfgrip.com
Second-order effects 83
www.pdfgrip.com
84 Internuclear spin-spin coupling
of the [AX]2 spin system, with the intense TV doublet and a series of
AJB patterns, but on account of the large value of 2/(31P31P), which
means that K, M, etc. are large, the central lines merge into a broad
central singlet and the outer lines vanish into the noise. The result is
a deceptively simple triplet. On the other hand, when two phosphorus
ligands are equivalent and are mutually cis9 2/(31P31P) is small for Pd11
and other complexes, K, M, etc. are small and now the lines crowd
ground the N doublet lines and only a doublet is observed, see Fig.
3.33, complex c. This behaviour has proved to be a valuable tool in
determining the stereochemistry of phosphorus complexes of Os11,
Ru11, Rh1, Ir1, Rhm, Irm, Pd11, Pt11, PdIV, and Pdiv.
Figure 3.35 shows the proton spectrum of [PtMe(PMe2Ph)3]+, already
encountered in Fig. 3.16, where the effect of the 31P atoms was
removed by double irradiation. The full spectrum contains a triplet at
8 1.68 due to the mutually trans PMe2 protons, a doublet at 8 1.29 due
to the PMe2 protons trans to the PtMe group, and a doublet of triplets
due to the Pt-Me group at 8 0.54. In each case, the signals are flanked
by 195Pt satellites.The methyl groups of the phosphine ligands cannot
in any way couple equally to the two phosphorus atoms and their
triplet is the deceptively simple triplet described above. Such triplets
are taken to indicate trans structures for such complexes, whereas
doublets arise from cis configurations.
An example of an organic molecule that has a four-spin [AX]2 spec-
trum is furan, whose proton spectrum is shown in Fig. 3.36 together
Ik. •>_/v/^-<sAA/LA_A_/v*^
\ I I I I I I I I I I I I \
1.8 1.6 1.4 1.2 1.0 0.8 0.6 0.4
1
H/ppm
Figure 3.35 The 400 MHz !H NMR spectrum of the complex [PtMe(PMe2Ph)3][PF6] in CDC13 showing
the three methyl resonances in the intensity ratio 4:2:1. The flanking resonances are 195Pt satellites. The
small triplet splittings of the signal at 8 1.68 are due to coupling to the two magnetically inequivalent but
chemically equivalent 31P atoms giving rise to a deceptively simple triplet. The Pt-Me at 8 0.55 is a doublet
of triplets due to coupling to the 31P atoms.
www.pdfgrip.com
Second-order effects 85
Figure 3.36 The 400 MHz 1H NMR spectrum of furan in CDC13. Coupling constants and splittings are
given in hertz. 1J(1H - 13C) = 201.4 Hz and '/OH - 13C) = 175.3 Hz respectively.
www.pdfgrip.com
86 Internuclear spin-spin coupling
3.6 QUESTIONS
3.1. Figure 3.6 shows the 1H spectrum of an ethyl group. Measure the
chemical shift of the two multiplets and their coupling constant.
3.2. Verify that the ratio between the coupling constants to 1H of the
nuclei 10B and n B in the [BH4]~ are in the ratio of the magnet-
ogyric ratios of the two boron isotopes (Fig. 3.8). What do you
expect the n B NMR spectrum of [BH4]~ to look like?
3.3. Figure 3.9 shows the proton spectrum of a CD2H group in iso-
topically substituted acetone. 2J(2H1W) =2.2 Hz. Calculate the
coupling constant 2/(1H1H) between the protons in the CH3
groups of normal acetone, ignoring isotope effects on /.
1
3.4. Doublet satellite signals are observed in the H spectrum of
acetone, close to the main singlet signal, with a spacing of 5.9 Hz,
which is the value of the coupling 2J(1H-13C). What will be the
pattern in the 13C spectrum of this carbonyl group, i.e. number
of lines and their relative intensities?
3.5. The 60 MHz !H spectrum of ascaridole features a tightly coupled
AB quartet at about 8 = 6.45 ppm. The line positions, starting at
the highest frequency one, are 395.5, 386.9, 385.5 and 376.9 Hz
from TMS. Calculate 3/(Ha-Hb) and the chemical shift between
Ha and Hb in Hz and ppm. Then calculate the relative line inten-
sities and positions expected when the same sample is observed
at 500 MHz. What is the exact chemical shift of the centre of the
quartet?
3.6. Explain why, in Fig. 3.25(a), the two halves of the AB quartet
have different intensities and so different linewidths.
3.7. Assign the spin systems, e.g. AB, to the following compounds,
a. ethyl chloride
b. CH2F2
H
c.
d.
e.
www.pdfgrip.com
Questions 87
f. CH2=CHBr
g. [PdCl2(PMe3)2]
h. /0c-[IrH3(PR3)3] (ignore any spins in the R groups).
3.8. Figure 3.37 shows the 19F resonance of the CF2H group of
1,1,1,2,2,3,3-heptafluoropropane, CF3CF2CF2H. The two fluorines
are equivalent and are coupled to all the other magnetically active
nuclei in the sample. Pick out the various multiplet patterns and
measure the coupling constants 2/(F-H), 3/(F-F) and 4/(F-F).
1
i ' i ' i '
40 20 0 -20 -40
Hz
Figure 3.37 The 19F resonance of the CF2H group of CF3CF2CF2H. The spec-
trum has been simulated using experimental coupling constants.
(b) 8(31P)
www.pdfgrip.com
Internuclear spin-spin coupling
100 Hz
3.10. In d8-THF, the 13C NMR spectrum with the effect of the protons
removed by double irradiation, of the lithiated carbon atom of
[(CH2=13CH)6Li] shows two sets of signals. Fig. 3.40 shows the
13
C NMR spectrum of the lithiated carbon atom of the major
isomer. The coupling is to 6Li. How many 6Li atoms are attached
to each carbon atom?
20 Hz
www.pdfgrip.com
Questions 89
The signal due to the lithiated carbon atom of the minor isomer
is shown in Fig. 3.41 at both -90°C when the molecule is rigid
and at -60°C when the atoms of the molecule are mobile and
the lithium moves between the vinylic CH carbon atoms in the
molecule. Use the coupling pattern to determine how many
lithium atoms are attached to a lithiated carbon atom in the rigid
structure, and how many lithium atoms visit each vinyl CH carbon
atom.
(a)
Figure 3.41 The 100.6 MHz 13C NMR spectrum of the lithiated carbon atom
of the minor isomer of [(CH2=13CH)6Li]n in d8-THF at (a) -90°c and (b)
-60°C. (Reproduced with permission from Brauer and Griesinger (1993) /.
Am. Chem. Soc., 115, 10871, copyright (1993) American Chemical Society.)
3.11. Figure 3.42 shows the 119Sn NMR spectrum of the mixture of iso-
topomers, [SnHnD3_J~, n = 0 to 3, although there is only a trace
of [SnD3]~. Identify all the signals present. Account for the multi-
plicity of each signal. Analyse the coupling patterns and indicate
which splittings are due to 1/(119Sn1H) and 1/(119Sn2H). Derive
the secondary isotope shift which occurs on the replacement of
1
H by 2H.
\ \ i i i i i i i
0 -5 -10
119
Sn/p.p.m.
Figure 3.42 The 134 MHz 119Sn NMR spectrum of the mixture of isotopomers,
[SnHnD3_J~, n = 0 to 3, in liquid ammonia at 20°C. (Reproduced with permis-
sion from Wasylishen and Burford (1987) /. Chem. Soc., Chem. Commun.,
1414.)
www.pdfgrip.com
90 Internuclear spin-spin coupling
3.12. Figure 3.43 shows the 100.62 MHz 13C NMR spectrum of the
mixture of isotopomers, CHnD3_nI, n = 0 to 3. Identify all the
signals present. Analyse the coupling patterns and derive
1 1
J( *C1H) and 1J(13C2H). Derive the secondary isotope shift which
occurs on the replacement of !H by 2H.
2 1 0 - 1 - 2
p.p.m.
Figure 3.43 The 100.62 MHz 13C NMR spectrum of the mixture of
isotopomers, CHnD3_nI, n = 0 to 3. (Reproduced with permission from
Sergeyev et al (1994) Chem.Phys.Lett., 221,385, copyright (1994), with permis-
sion from Elsevier Science.)
Figure 3.44 Expansions of the 400 MHz *H NMR signals from the hydride
atoms of [Ru(CO)H2(PPh3)3] in CD3C6D5.
www.pdfgrip.com
Questions 91
3.14. Figure 3.45 shows the 1H NMR signals for the CH2 protons of
frans-[PdCl2{P(CH2Ph)2Ph}2]. Account for the appearance of the
spectrum. Predict what signal would be observed for the CH2
protons of fr0ws-[PdCl2{P(CH2Ph)Ph2}2].
Figure 3.45 The 220 MHz 1H NMR spectrum of the CH2 protons of trans-
[PdCl2{P(CH2Ph)2Ph}2]. (Reproduced with permission from Nelson and
Redfield (1973) Inorg. Nucl Chem. Letts., 9, 807, copyright (1973) Elsevier
Science.)
3.15. Figure 3.46 shows the 6Li and 15N NMR spectra of the product
obtained when Me215NCH2CH215NMe2 is added to a solution of
6
LiBun in d8-toluene. Suggest a structure for the product. Account
for the multiplicity of each signal.
(a)
Figure 3.46 44.15 MHz 6Li and 30.408 MHz 15N NMR spectra of the product
obtained when Me215NCH2CH215NMe2 is added to a solution of 6LiBun in d8-
toluene. (Reproduced with permission from Nicholls et al. (1997) /. Am. Chem.
Soc., 119, 5479, copyright (1997) American Chemical Society.)
www.pdfgrip.com
92 Internuclear spin-spin coupling
3.16. Figure 3.47 shows the 235.36 MHz 19F NMR spectrum of cis-
[Ag(CF3)2Cl(CN)]- in THF. Account for the multiplicity of each
signal.
3.17. Assign the following proton signals from menthol, Fig. 3.2, using
the multiplicities and coupling constants.
8 2.18, septet (7 Hz) of doublets (3 Hz).
8 1.12, doublet (12 Hz) of doublets (10 Hz) of triplets (3 Hz).
SPECTRAL INTERPRETATION
www.pdfgrip.com
Questions 93
12 9 6 , 3 0 - 3
1
H/ppm
frequency, often below 10 ppm, and that metal alkyls are found to low
frequency around and above TMS. Hydrogen bonded to metals occurs
in a wide range also at low frequency of TMS and varies from
about -3 ppm, e.g. [HRe(PPh3)3(CO)2], to -50 ppm, e.g. [HIrCl2(phos-
phine)2], though there are one or two exceptions with shifts to high
frequency of TMS. The second piece of information is the integral
trace of the spectrum, which gives the area under each resonance and
www.pdfgrip.com
94 Internuclear spin-spin coupling
1 2 1 11 0 9 8 7 6 5 4 3 2 10
1
H/ppm
Figure 3.49 The 400 MHz 1H NMR spectrum of C3H5O2Br in CDC13.
www.pdfgrip.com
Questions 95
8 7 6 5 4 3 2 1 0
1
H/ppm
Exercise 1 The 60 MHz !H NMR spectrum of C3H4SO2 in CDC13. Deduce a
structure which is consistent with this spectrum. (Reproduced with permis-
sion of Varian, Inc.)
Possible answers: CH3SO2OCH, HOCH2SOC=CH, H H
C=C
C-SO 2
H2
www.pdfgrip.com
96 Internuclear spin-spin coupling
group must then take the other. The molecule is as shown in the
margin.
Now apply the same approach to Exercises 1 to 3. Bear in mind the
following: (a) The fluorine resonances of the compounds containing
fluorine are not visible but the proton resonances may be coupled to
the 19F and so split into multiplets. The same comments apply also
to the phosphorus compound, (b) Alcoholic protons are variable in
position because of differing hydrogen bonding effects. The numbers
on the spectra give the relative numbers of protons contributing to
each resonance and have been obtained from integral traces. Exercises
1 and 2 show expanded regions of the spectra so as to allow fine struc-
ture to be distinguished. The answers are given below, though several
structures are given, only one being correct. All the features of a spec-
trum should be explicable from the structure. It is instructive to also
consider what differences in the spectrum would be obtained from the
incorrect structures given.
A spectrum is shown in Fig. 3.50 with numbers corresponding to
the integrals of each group of resonances. The empirical formula is
C6H14O2. Evidently there are five types of hydrogen and all show
spin-spin coupling. The doublet and triplet at 1.37 and 1.25 ppm are
in the chemical shift region typical of CH3R and the integrals indicate
that they are due to a CH3 and two identical CH3 groups. The former
are coupled to a single proton which should have a quartet resonance.
This is seen at 4.7 ppm, the line spacings being equal, and the shift
1 4.0 3.5
LA.
8 7 6 5 4 3 2 1 0
1
H/ppm
Exercise 2 The 400 MHz J H NMR spectrum of C2H3F3O in CDC13. Draw a structure which is consistent
with this NMR spectrum.
Possible answers: CF2HCHFOH, CF3CH2OH, CH2FCF2OH.
www.pdfgrip.com
Questions 97
C3F4H40
^*v**~~jJ*»~Jt^^ L/
8 7 6 5 4 3 2 1 0
1
H/ppm
Exercise 3 The 60 MHz *H NMR spectrum of C3F4H4O in CDC13. Deduce a
structure which is consistent with this spectrum. (Reproduced with permis-
sion of Varian, Inc.)
Possible answers: CF2HCH2CF2OH, CF3CHFCH2OH, CF2HCF2CH2OH,
CH3CF2CF2OH.
TMS
www.pdfgrip.com
98 Internuclear spin-spin coupling
IMS
II Tb ill
CeHeIHA,
I III
Jil
8 7 6
7.5
5 4
2
A
7.0
3
6.6
2 1 0
1
H/ppm
Figure 3.51 The 400 MHz !H NMR spectrum of 2-nitroaniline in CDC13. The intensities of the signals are
given above each one.
alent ethyl groups CH3CH2-. This accounts for all the atoms in the
molecule except the two oxygen atoms. Because the methyl groups
are identical, the oxygen atoms have to be distributed symmetrically
and this leads inevitably to the structure CH3CH(OCH2CH3)2. The
CH2 pairs of hydrogen atoms in this molecule are diastereotopic and
so non-equivalent which accounts for their complex spectrum.
In Fig. 3.51, we see four groups of resonances in the aromatic/alkenic
region of shift which exhibit a larger and a smaller coupling. There
are also two uncoupled protons in the region typical of RNH and, in
view of the empirical formula given on the figure, this suggests an NH2
group which leaves us with a C6H4 aromatic nucleus and an NO2 group.
We thus have a nitro-aniline and need only to deduce which isomer
we have. This can be done using the coupling patterns. Were it the
para isomer, the two pairs of protons would be equivalent and an
[AB]2 pattern with two groups of resonances would result. Were it the
meta isomer then there would be an isolated hydrogen (H^ which
would show only the small coupling. This is plainly not the case.
However, in the ortho isomer all protons have near neighbours and
7.6 7.4 7.2 7.0 6.8 6.6 6.4 6.2 6.0 5.8 5.6
Figure 3.52 A 400 MHz !H NMR spectrum and its integral of 2,6-dichlorophenol in CDC13. The intensi-
ties of the signals are given above each one.
www.pdfgrip.com
Questions 99
www.pdfgrip.com
www.pdfgrip.com
Nuclear magnetic
relaxation 4
4.1 RELAXATION PROCESSES IN ASSEMBLIES OF
NUCLEAR SPINS
www.pdfgrip.com
102 Nuclear magnetic relaxation
the xy plane. It turns out that this system can be perturbed in two ways,
and that we have to expect that there may be two relaxation processes
with different relaxation times, which are called Tl and T2, or rates of
relaxation ^ and R2. We have already seen in Chapter 1 (Fig. 1.6) that
a Bl radiofrequency pulse can swing the total nuclear magnetization
away from its equilibrium position in the z direction, and this is essen-
tially a perturbation of the system. If we apply a rather long pulse, we
can swing the magnetization back into the z direction but pointing in
the opposite direction. Such a pulse is called a 180° pulse, the reason
for this name being evident in Fig. 4.1. The magnetization has been
inverted and, immediately following the end of the pulse, relaxation
processes start to return the magnetization to its normal state. Thus the
magnetization decays from -Mz via zero to Mz via first-order rate
processes. The characteristic time for this process is 7\. The process is
also called longitudinal relaxation since it takes place in the direction
of U0, and it is also called spin-lattice relaxation. In all cases it must be
emphasized that the inverted magnetization has higher energy than the
normal magnetization and that the return to equilibrium involves an
exchange of magnetic energy with the surroundings.
If instead we use a 90° pulse to perturb the spin system, we move
the magnetization into the xy plane as in Fig. 4.2. Now the magneti-
zation in the z (BQ) direction is zero, and this returns to its normal
MzA
// 180° pulse //
~-^ ^-Mz
relaxation relaxation
BO BO
i \^ relaxation /^
-^ ' ^^
Figure 4.1 The Tl relaxation process. If the magnetization is inverted, then
it has to return to its equilibrium state and does so by decaying to zero and
then increasing again in the normal B0 direction. This process involves an
exchange of energy between the spins and their environment.
www.pdfgrip.com
Relaxation processes in assemblies of nuclear spins 103
Figure 4.2 The T2 relaxation process. A 90° pulse swings the magnetization
into the xy plane around the Bl vector. This is shown as stationary in the
figure as if the observer were rotating in the same direction at the Larmor
frequency, thus giving a static picture. There is a spread of nuclear frequen-
cies, which causes the spins to fan out and reduce the resultant Mxy. Mz
increases at the same time due to the 7\ process.
www.pdfgrip.com
104 Nuclear magnetic relaxation
Wm - -^ (4.1)
This is true for a singlet in a perfectly homogeneous magnetic field,
but in practice there are often extra contributions to the linewidth due
to inhomogeneity of the magnetic field and unresolved coupling.
The nuclei of atoms are extremely well isolated from their surround-
ings and, because the energy of NMR transitions is small, the chance
of a spin transition occurring spontaneously is negligibly small. The
fact that relaxation times can be quite short indicates that transitions
are stimulated, and we must thus consider the various ways that this
can happen.
We have already noted in section 2.1 that the magnetic field at a given
nucleus due to the magnetic moment of a near-neighbour nucleus is
very high but is averaged to zero by the random rotational diffusion
of the molecule in which the nuclei reside. The magnitude of this field
is such that the instantaneous chemical shift displacement of one 1H
nucleus due to the other in a methylene group can be as high as
150 000 Hz. As the group rotates, this field varies by such an amount
on each side of zero. Thus the nuclei have instantaneously different
precession frequencies since all possible orientations of the molecules
will exist at any one instant. Randomization of the frequencies means
not that all will have the same frequency in the long term but rather
that once out of step a nucleus is just as probable to move further
away from its companion's frequency as to reconverge to it. This
dipole-dipole fluctuating field then is the cause of the loss of coher-
ence between spins and so the source of the T2 relaxation process.
The chaotic random motion of a solute in a solvent is called
Brownian motion. This has a timescale that depends upon a number
of factors such as mass of solute, solution viscosity and temperature.
Because the motion is random, this timescale is characterized by a
somewhat loosely defined term, the rotational correlation time TC. This
is the time taken on average for a solute molecule to rotate by one
radian or, more precisely, the time interval after which the molecular
motion contains no vestige of its earlier angular momentum, i.e. has
lost all memory of its previous behaviour. Not only the overall rota-
tion of the molecule contributes to TC, but also local motion, such as
the rotation of a methyl group. The time TC is typically 10"11 s for small
molecules in liquids of low viscosity, which converts on the frequency
scale to 1011 rad s'1 or 15 920 MHz. This is around the maximum rate
of motion in the system, and all slower rates of motion can exist. The
frequency spectrum of such random motion and associated magnetic
www.pdfgrip.com
Dipole-dipole relaxation 105
i i i i i i
1 10 100 1000 10000 100000 1000000
MHz
Figure 4.3 Intensity of fluctuations in magnetic fields in a liquid sample due to Brownian motion, as a
function of frequency, (a) TC - lO'11 s-1. (b) TC - 10;10 s-1. (c) TC = 10'9 s'1. (d) TC - 10'8 s-1. Note that a
different vertical scale is used for each plot as the intensity of fluctuations9 in1 the flat region of the plot
is proportional to TC. TC ~ 10~n s-1 is found for very small molecules, TC ~ 10~ s' is commonly encountered
with molecules of molecular mass, 1000 to 3000 D, while TC > 10~9 s'1 is normally found for molecules of
molecular mass > 5000 D.
www.pdfgrip.com
106 Nuclear magnetic relaxation
4-TTfl3 Tl
Tc =
~3F"r
where T\ is the viscosity of the liquid, T is the temperature and a is
the radius of the molecule. Thus if we vary the viscosity of a sample,
or its temperature, or the mass of the solute molecule, we will change
TC. Examination of equation (4.2) shows that, when 4TT2v2Tc2« 1, the
intensity of the relaxation field is proportional to TC. For the higher
NMR frequencies and the longer correlation times, we start to leave
the flat portion of the noise spectrum and the relaxation field starts
to decrease. Provided we limit our range to the portion of these curves
where 4ir2v2Tc2 « 1, then the relaxation field and so the relaxation rate
increases with TC. Long relaxation times thus occur for low viscosity,
high temperature and small molecular mass.
It will be seen from equation (4.1) that the relaxation field intensity
has its flat frequency response when the quantity 4ir2v2Tc2 is very much
less than unity, i.e.
4^2V2<< _1_
www.pdfgrip.com
Dipole-dipole relaxation 107
For / = 1/2 nuclei of the same isotope situated in the same mol-
ecule, the intramolecular dipole-dipole relaxation rates for the yth type
of nuclei, being relaxed by n different types of nuclei of the same
isotope are
1 4Tc
-J—-I? -2WM ^ i ^ V !
R +
T1DD - ™ - Ml + co V IT^oV ) J£ jf
£-^-^rfVr£v)^Ji
where a = 3|m,02/j2/320'Tr2, |JLO is the permeability of a vacuum, r is the
distance between the nuclei, y is their magnetogyric ratio, to is the
NMR frequency in rad s'1, and h is (Planck's constant)/27T. The way
these rates of relaxation vary with TC is shown in Fig. 4.4 for two
protons at two different spectrometer frequencies, to. The main feature
to note is the Tl minimum, which marks the limit of the extreme
narrowing region and the way this moves with spectrometer frequency.
The higher this frequency, the shorter becomes the maximum
permitted TC, the result being that, for large complex molecules where
the highest frequencies may be needed to give the necessary degree
of resolution, the increased correlation times of such molecules may
result in reduced T2 and so increased linewidths. The expression for
riDD and r2DD is much simpler in the extreme narrowing region
TIDD.
10-i
1-
extreme
narrowing
0.1 -
0.01 -
0.001 -
0.0001 -
0.00001
10-12 10-11 10-10 10-9 10-8 10-7 10-6
TC (s)
Figure 4.4 Variation of T1DD and T2DD with TC for two different spectrometer
frequencies. The figures given apply to two protons separated by 160 pm.
www.pdfgrip.com
108 Nuclear magnetic relaxation
1 1 ft 1
T =T =tflDD=tf2DD= 10«74Tc 2 n (4.3)
^ 1DD ^ 2DD /= 1 (/ =£ /)r//
Two further points should be emphasized. In the first place, the rate
of dipole-dipole relaxation depends upon the fourth power of the
magnetogyric ratio, and nuclei with large magnetic moments will be
most strongly subject to such relaxation, for example, 1H or 19F. The
mechanism will be of lesser importance for nuclei with smaller magne-
togyric ratios. Secondly, the efficiency of relaxation depends strongly
upon the inverse distance between the spins. The effect of more distant
spins will be almost negligible.
For an / = 1/2 nucleus in the same molecule as n different / = 1/2
nuclei with magnetogyric ratios yl and -ys and frequencies o)j and oos
rad s"1, respectively, the intramolecular dipole-dipole relaxation rates
of one (species I) due to interaction with other (species S) are
2 2Tc 6Tc
T-
*1DD
= RmD = ^i
*
rfi. f ,2 2 +
[1 + (&>!-0) )^T 2
S C l+Wl2T(2
1?T 1 " 1
+ ^5 ^ J_
l + (coI + cos)VL=1^,)'"/
7
2DD ^
^ = *2DD ^ +3f^sf
\ 1 + <0S Tc2y.=^^..6
c +I ^) $ ^
These equations, though more complex, give plots of form very
similar to those depicted in Fig. 4.4. In the extreme narrowing limit
www.pdfgrip.com
Quadrupolar relaxation 109
www.pdfgrip.com
110 Nuclear magnetic relaxation
oblate prolate
Vzz is chosen to be the largest component of the EFG tensor, and this
can then be described by two quantities, Vzz and i\. If the system is
axially symmetric so that Vyy = Vxx then TI = 0. The Vxx, etc. are calcu-
lated as the sum of contributions from all charges i
VKX = Sfcr-W - r,2) (4.6)
where ri is the distance of charge qi from the nucleus and xt is the
coordinate of the charge in the axis system chosen. Similar expres-
sions give V and Vzz.
The equation describing the quadrupole relaxation times (TIQ, T2Q)
of a quadrupolar nucleus with spin / situated in a molecule with an
isotropic correlation time TC and in the extreme narrowing region is
www.pdfgrip.com
Quadrupolar relaxation 111
tt^-zmftM+i^
where Q is the quadrupole moment of the nucleus of spin, /, e is the
(4.7)
3 5 7 9
/ 1 2 2 3 2 2
4 _8 1 _2Q 2
F(I) 5 3 25 5 147 27
www.pdfgrip.com
112 Nuclear magnetic relaxation
EFG with sufficient accuracy. Equation (4.4) shows that the EFG is
proportional to the inverse cube of the distance of the charge from
the nucleus. Thus charge close to the nucleus has the predominating
effect, and it is here that the calculations are least accurate. In fact,
there is some confusion in the literature on this point. In the solid
state, the EFG arises from quite distant charges as well as those close
by, in the same way that a Madelung constant is calculated. In a liquid,
the movement of the molecules reduces the distant effects to zero and
the EFG arises quite locally around the nucleus. For instance, in the
anion [A1C13(NCS)]- both the quadrupolar nuclei 27A1 and 14N have
long relaxation times and show spin-spin coupling. There are two
points of low EFG in the molecule, and this can only arise if the EFG
arises in the region quite close to the nucleus. The question is, of
course, how close has the charge to be, to be effective?
The cases of greatest interest are those in which the relaxation times
are relatively long, since this gives the best resolution of resonances
and the possibility of seeing coupling effects. This means that we
should understand the requirements for obtaining a low EFG at a
nucleus. We can do this easily if we remember that for a traceless
EFG tensor the sum of the diagonal elements is zero
If the system is axially symmetric (i.e. T] = 0), Vyy = Vxx, and it follows
that if Vzz is zero then all three terms must be zero and the EFG
vanishes. Thus in an axially symmetric system we need only calculate
Vzz. We take for the model in Fig. 4.6 a tripod of bonds to a nucleus
N disposed regularly around the z axis with effective, equal charges q
called ql9 q2 and q3 at a distance r from N.
www.pdfgrip.com
Quadrupolar relaxation 113
www.pdfgrip.com
114 Nuclear magnetic relaxation
Table
59
4.2 The predicted quadrupolar electric field gradients and observed
Co linewidths for some cobalt(III) amine complexes
Structure Relative magnitude of Complex Linewidth (Hi)
quadrupolar electric
field gradient
MA6 0 [Co(NH3)6]3+ 2 87
MA5B 2 [Co(NH3)5(N3)] + 350
d5'-MA4B2 2 ds-[Co(NH3)4(N3)2]++ 370
trans-MA4B2 4 frans-[Co(NH3)4(N3)2] 520
/flc-MA3B3 0 /ac-[Co(NH3)3(N3)3] 300
raer-MA3B3 3 mer-[Co(NH3)3(N3)3] 1240
www.pdfgrip.com
Quadrupolar relaxation 115
www.pdfgrip.com
116 Nuclear magnetic relaxation
www.pdfgrip.com
Spin rotation: detailed molecular motion 117
(a)-2 - 1 0 1 2 (e)-2 -1 0 1 2
i
(b)-2 - 1 0 1 2 (f) -2 -1 0 1 2
(c)-2 - 1 0 1 2 (g)-2 -1 0 1 2
(d)-2 -1 0 1 2 (h)-2 -1 0 1 2
Figure 4.8 The resonance line shape of a spin-1/2 nucleus coupled to a spin-
1 nucleus having various rates of quadrupole relaxation, (a) Tl ~ 0.05/Tr/. (b)
T; - 0.1/7T/. (c) TI « 0.2/ir/. (d) 7\ « 0.4/7T/. (e) 7\ - 0.8/ir/. (f) Tl « 2.0/ir/. (g)
Tl ~ 4.0/ir/. (h) 7\~ 8.0/TT/. The intensities are not to scale. The horizontal
axis is in units of Av//.
www.pdfgrip.com
118 Nuclear magnetic relaxation
www.pdfgrip.com
Spin rotation: detailed molecular motion 119
Temperature /K
80 90 100 120 140
30
20
CD
.i 10
4-»
I 6
X
05 4
A
o
a:
.J
12 10 8
3
10 /Temperature (K)
Figure 4.9 Measured values of the spin-lattice relaxation time, 7\, for liquid
14
N2 (in ms), o, and liquid 15N2 (in s), x, on the liquid vapour coexistence line
versus temperature (on a scale of 103 K/71, where T is the absolute tempera-
ture); c.p. indicates the critical point, and b.p. the boiling point at 1 aim.
(From Powles et al (1975) Mol Phys. 29, 539, reprinted with permission,
copyright (1975) Taylor & Francis.)
www.pdfgrip.com
120 Nuclear magnetic relaxation
Temperature/°C
10080 60 40 20 0
100
80
60
40
20
.0)
10
8
6
4
3.0 3.5
3
10
-y- /K
www.pdfgrip.com
Scalar relaxation 121
1 _2 7I 2 /? 0 2 (CT||-C7 J .) 2 T C
T
1
1CSA
1S
LJ
www.pdfgrip.com
122 Nuclear magnetic relaxation
8.48 T
1.5- [Pt(CN)4f-
1.0-
0.5-
4.70 T
[Pt(CN)4f-
0.0-
-0.5-
[Pt(CN)6f-
Figure 4.11 The 195Pt relaxation rates for aqueous solutions of K2Pt(CN)4 and
K2?t(CN)6 at B0 = 4.70 and 8.48 T as functions of inverse temperature. The
i95pt relaxation rates of K2Pt(CN)6 are independent of BQ. (From Wasylishen
and Britten (1988) Magn. Reson. Chem., 26, 1075, copyright John Wiley and
Sons Ltd, New York, reprinted with permission.)
www.pdfgrip.com
Scalar relaxation 123
400 MHz
250 MHz
90 MHz
www.pdfgrip.com
124 Nuclear magnetic relaxation
(b)
Figure 4.13 The 13C NMR spectra of [Mn(CH2Ph)(CO)5] in CH2C12 at (a)
30°C, and (b) -87°C. Note the improvement in both the linewidth and
signal:noise at low temperature. (Reproduced from Todd and Wilkinson
(1974) /. Organomet. Chem., 80, C31, copyright (1974), with permission from
Elsevier Science.)
www.pdfgrip.com
Examples of 13C relaxation times 125
rotation and has the shortest 7\ value of the CH carbon atoms. The
ortho- and meta-CH and the vinylic CH groups lie off the axis and
the hydrogen atoms move with respect to the carbon atoms due to the
rotation. The result is that TC is shortened and Tl becomes longer.
Thiamine hydrochloride provides a more representative example.
The 7\ values are given in Fig 4.14 for the compound in CH3OH and
CD3OD. Note that certain non-proton bearing carbon atoms have
much longer 7\ values than the proton bearing ones, but the Tl is
reduced for these non-proton bearing carbons in the protio solvent
due to the proximity of the solvent protons. The Tl of the methyl
group is lengthened by rotation relative to the CH carbon atoms. By
(a) (b)
13
Figure 4.14 The C relaxation times of thiamine hydrochloride in (a) CH3OH and (b) CD3OD, measured
at 25.15 MHz.
(4.2)
www.pdfgrip.com
126 Nuclear magnetic relaxation
4.8 QUESTIONS
www.pdfgrip.com
Questions 127
5.2s 5.5s
H H H H
C—C C—C
// \\ 82s J 36s / \
a p
1.1s HC C —C =C -C=C—C x , CH
\ / \\ //
c=c c—c
H H H H
(b) When the 13C Tl values are determined at 63.1 MHz, values
of 15 s and 30 s were found for Ca and Cp. Explain why the
Tl values have decreased by so much.
4.9. The 13C Tl values have been determined at 25.15 MHz for the
ipso carbon in bromobenzene as 3.6 s for C6H579Br and 16 s for
C6H581Br. Explain why the two values are different. What is going
to be the effect on these Tl values of increasing the magnetic
field so that the 13C frequency becomes 50.3 MHz?
4.10. Although the relaxation of proton bearing 13C nuclei is normally
dominated by the dipole-dipole relaxation mechanism, no signif-
icant dipole-dipole relaxation has been found for 103Rh in
rhodium hydrides. Explain this observation.
4.11. In Fig. 4.11, it is shown that for J?0 - 8.48 T and Iffi/T = 2.85, the
195
Pt relaxation rate in [Pt(CN)4]2- is given by In R± = 0.8.
Calculate the 195Pt linewidth of this anion under these conditions
assuming Tl = T2 and a perfectly homogeneous magnetic field.
The measured value of the linewidth is, in fact, 90 Hz. Calculate
T2 and estimate the contribution of scalar relaxation to this value.
www.pdfgrip.com
www.pdfgrip.com
The spectrometer 5
We have seen in the last three chapters that there are three principal
parameters in an NMR spectrum: the chemical shift, the coupling
and, less evidently, the relaxation. The present chapter is devoted to
showing in more detail how the spectrum is obtained and how these
parameters may be derived from the data collected.
Chapter 1 discussed how a 90° B1 pulse at the nuclear frequency
can swing the magnetization from the direction of the magnetic field
into the xy plane, and showed that this magnetization, which is now
rotating at the nuclear frequency, can be detected in a suitable coil.
In Chapter 4 we discussed how this magnetization is affected by a
variety of relaxation processes, in particular, how T2 leads to an expo-
nential reduction with time of the xy magnetization. The spectrometer
output thus consists of a signal at the nuclear frequency, which decays
with time until it is no longer detectable. We need to know how
this can be translated into typical spectral form. The equations
governing the behaviour of the transverse and longitudinal magneti-
zation M and Mz and their return to equilibrium following a 90° Bl
pulse are
(M,)t = (M z )Jl-exp(-^-)|
L \ M/J
i.e. the transverse magnetization falls from its maximum value (equal
to Mz) to zero after sufficient time has elapsed. This behaviour is
summarized diagrammatically in Figs 4.1 and 4.2.
In general, the rate of decay of the spectrometer output is faster
than predicted from the value of T2 in a given system. This occurs
because of inhomogeneities in the magnetic field throughout the
sample due principally to imperfections in the magnet system, which
means that the nuclear frequencies are slightly different in different
parts of the sample volume and this increases the rate at which the
spins throughout the sample lose phase and coherence. We write that
the apparent relaxation time, T2*, is equal to
www.pdfgrip.com
130 The spectrometer
J___JL 1
T
1 # ~~ 1T LT
2 2 inhomo
! TT
^--[Digital Filters]
.u..r -] Computer
i | [Data memory I
! i i
Interactive! [Printer] |Disk storage]
display [
www.pdfgrip.com
The magnet and field homogeneity 131
itself is then cooled with liquid nitrogen, which in turn is also protected
by the vacuum jacket (Fig. 5.2). Modern systems are very efficient with
25 1 of liquid helium lasting for at least three months. The liquid
nitrogen needs topping up regularly, typically once a week. Once the
current has been established in the superconducting coil, it will
continue to give an acceptable magnetic field for at least ten years,
before the current has decayed sufficiently to need to be topped up
again. The current in the superconducting coil is quantized, as is the
magnetic field. If an NMR tube containing CHC13 is placed in the
magnet, and a single pulse spectrum taken every hour, then it will be
found that the position of the signal remains constant for a while, and
then jumps one or two hertz to lower frequency, and again remains
constant for a while. This experiment monitors the quantized jumps
of the magnetic field, which of course have to be counteracted to give
a permanently stable field.
Although there are no reasons to believe that a static magnetic field
can cause harm to personnel, it is usual to mark a 5 gauss (0.0005 T)
contour line around the magnet, and to try to stay outside this region
as far as is possible. The major recognized hazard associated with the
stray magnetic field is its ability to capture large metallic objects such
as tools and gas cylinders which accelerate rather rapidly towards the
magnet! Nor should those with heart pacemakers go anywhere near.
Anything, including floppy discs and credit cards, containing magnetic
records can be damaged by the stray magnetic field.
www.pdfgrip.com
132 The spectrometer
5.1.2 Shimming
Two sets of coils are wound to produce magnetic field gradients to
cancel those inherent in the main magnetic field. The first set consists
of superconducting coils, which are wound within the liquid helium
bath and are adjusted as part of the magnet installation. The currents
in these coils are not subsequently changed by the operator, but the
operator has to be aware of their existence as they can quench, i.e.
lose their current, and the resolution deteriorate. The second set of
shim coils is mounted around the probe and the currents in them are
adjusted to cancel any remaining field gradients and give optimum
resolution. This second set of coils comprises:
1. Zero-order coil. There is a single zero-order coil with its axis
along the direction of the superconducting solenoid and this is
used to adjust the main magnetic field strength within narrow
limits.
2. First-order coils. There are three first-order coils called XI, Yl, and
Zl, which produce magnetic fields as the functions x, y, and z. They
produce field gradients shaped like p-orbitals. The x and y coils are
aligned at right angles to the main magnetic field, while the z coil
is coaxial with the main field.
3. Second-order coils. There are five second-order coils called XY,
X1-Y1, ZX. ZY, and Z2, which produce magnetic fields as the func-
tions Jty, x2 - y2, zx. zy, and z2. They produce field gradients shaped
like d-orbitals.
www.pdfgrip.com
The magnet and field homogeneity 133
4. Third-order coils. There are seven third-order coils called Jt3, Y3,
Z2X. Z 2 Y, ZXY, Z(X2-Y2), and Z3, which produce magnetic
fields as the functions x3, y3, z2x. z2y, zxy, z(x2-y2), and z3. They
produce field gradients shaped like /-orbitals.
5. The magnets designed for the highest magnetic fields are often also
equipped with fourth- and fifth-order z-shims.
It has to be emphasized that the quality of the spectra depends
absolutely on minimizing the field gradients with the shims. Adjusting
all the shims, however, is a highly skilled operation and beyond the
scope of this book, but some excellent texts are available, and listed
in the Bibliography. The inexperienced user is best advised to restrict
changes in the shim currents to those in the Z and Z2 coils when the
sample is spinning. The shim currents are adjusted by optimizing
the height of the lock signal for maximum, see section 5.3. This may
need to be accompanied by adjustment of the lock phase for maximum J
height. If the X and/or Y shim currents are mis-set, then the magnetic
field at a given point in the sample fluctuates as the sample spins, the 20 0 -20 Hz
nuclear frequencies are modulated at the spinning frequency and the
NMR signals will be flanked by spinning side-bands (Fig. 5.3). These Figure 5.3 A 1H NMR
shim currents then have to be adjusted with a non-spinning sample. spectrum of CHC13 show-
Shimming is critically important in obtaining a high-quality NMR spec- ing spinning side bands
due to mis-setting of the
trum. A good quality control of the shim settings is to routinely current in the X shim
examine the signal from the TMS reference. It should be a singlet with coil. The sample tube was
well-resolved 29Si satellites, 2J(29SilH) = 6.6 Hz. It is also useful to spun at 20 Hz and the
watch the FID as it accumulates. It is often possible to identify a spinning side bands are
symmetrically situated at
sample that is giving poor resolution after only one FID has been ± 20 Hz of the main signal.
acquired. For 1H, the FID should last for several seconds, and the (Reproduced with per-
presence of a beat after one or two seconds while it may be due to mission from Braun et al
coupling, often indicates poor resolution. (1996) 100 and More Basic
The position of the bottom of the NMR tube and the top of the NMR Experiments, VCH,
Weinheim.)
solution in the tube distort the magnetic field and so contribute to
the field gradient. Provided a gauge is used to position the spinner
on the NMR tube, so that the bottom of the NMR tube is always in
the same position, and a relatively long sample length is used, then,
on change of sample, it should only be necessary to adjust the shim
currents in the Z and Z2 coils when the sample is spinning.
Manual shimming is becoming replaced by computer shimming. For
many years, it has been possible to use iterative routines to adjust the
shims automatically, but these routines are slow, and are normally only
used to compensate for drift during long data accumulations. It is
becoming common for probes to be equipped with a z-magnetic field
gradient coil, see section 5.8.6. This coil can be used to map magnetic
field inhomogeneity in the z-direction, and the Z, Z2, Z3, Z4 and Z5
corrections are then calculated and applied by the computer. This only
takes a few minutes and uses the same methods as will be described
in Chapter 11 for magnetic resonance imaging. Less common is for
probes to be equipped with #-, y- and z-magnetic field gradient coils,
www.pdfgrip.com
134 The spectrometer
www.pdfgrip.com
Field-frequency lock 135
gasket —
outer inner glass
glass tube" tube
decoupling receiver
coil coil
Figure 5.4 A cross-section through a typical NMR probehead. Note all the
glass in the probehead which makes it fragile. The coils are wound as saddles
on the glass tubes (Helmholtz double coils) to give the RF field normal to
B0. (Reproduced with permission of Bruker Spectrospin.)
the probehead means that it is fragile and failure to insert the NMR
tube on a cushion of air can result in breakage.
www.pdfgrip.com
136 The spectrometer
Stabilizer
output
www.pdfgrip.com
The transmitter 137
1 Dispersion
— PSD >-
I 1 to magnet
field control
, , I Receiver | I 90° Phase
~~~1^-^" switch I Chan9er
Probe ^T '
I ^^^--^ Transmitter |
1
switch I 1 Absorption
' I PSD ^~
pulse J^ck 1 ^ to monitor
frequency] [frequency screen
Frequencyl L Phaqp I
Synthesizer Mhase
i— 1 ' [adjustment
Computer
Figure 5.7 A block diagram of the field-frequency lock system. The frequency
synthesizer provides both the pulsing frequency and the lock frequency. The
pulsing produces sidebands of the lock frequency and the lock operates on
one of these. The continuous output of the nuclear lock signal is fed to two
phase sensitive detectors (PSD, to be described in section 5.5.1) operating in
quadrature which produce the dispersion and absorption signals to correct the
magnetic field and enable the signal to be monitored. Means are provided to
adjust the phase to ensure that the spectrometer is operating at its optimum
position.
The transmitter provides the pulses to stimulate the FID and also to
manipulate the nuclear spins in a variety of ways. The centre of the
transmitter system is a quartz controlled frequency synthesizer which
generates the frequencies needed for a given experiment and also that
required for the lock system described above. The frequencies
produced are locked together so that they always have the same rela-
tionship and are continuous and coherent, which is to say that they
are pure, continuous sine waves with no irregularities. Some of the
outputs are also used in the receiver, as we shall see. The transmitter
frequencies are used to drive the power amplifiers which produce the
pulses but have first to be processed. Pulses are produced by a switch
www.pdfgrip.com
138 The spectrometer
which applies the signal to the amplifier for the required short time.
Such pulses may be rectangular or they may be shaped by a computer
controlled device to give pulses with either a broad bandwidth or a
narrow bandwidth. These pulses are effectively bits of the input contin-
uous wave. Means are also provided to shift the pulse oscillations rela-
tive to the drive signal, in other words to phase shift them. The
processed signals are then amplified in the power amplifiers, one for
each frequency, and provision is made for the output to be varied in
intensity, the degree of attenuation usually being expressed in a dB
(decibel) scale, the more dB, the less the intensity. The decibel scale
is a log scale, being related to power, P, by
dB = 101og]0(^) = 201og10(^)
\r2' \v 2f
where Pl and P2 are the powers and Vl andV2 are the applied voltages
being compared. Hence increasing the dB value by 3 units approxi-
mately halves the power being used. However, as the radiofrequency
field strength is proportional to voltage, to halve it, an approximate
increase in the dB value of 6 units is required.
It is particularly important to be able to control the relative phases
of the output signals. This can be understood by referring to Fig. 5.8.
The pulse gives the field Bl in the xy plane, rotating with the nuclei
on, say, the x' axis in the rotating plane and the nuclear magnetiza-
tion rotates around B^ into the yf axis. If we use a 90° pulse then this
is called a (90°)^ pulse. If we phase shift the transmitter output by 90°
(do not confuse pulse lengths with phase shifts even though both are
expressed in the same units), this moves Bl 90° in the x'y' plane to
the y' axis and the nuclear magnetization now rotates into the -x' axis.
This is known as a (90°)^ pulse. It is equally possible to place the
magnetization along the jc' or -y' axes using a (90°)^ or (90°)_r phase
shifted pulse. We shall find this facility extremely useful below.
Figure 5.8 The effect of the phase of the 90° pulse on the magnetization with
respect to the x'y' rotating frame. The direction of the arrow within the circle
represents the direction of the magnetization in the x'y' plane after applying the
90° pulse about the jc, y, -Jt, or -y axes as indicated by the subscripts after the
bracket.
www.pdfgrip.com
The detection system 139
www.pdfgrip.com
140 The spectrometer
are exactly the same, then the nuclear signal is rectified with, say, the
negative half cycles reversed in polarity, and a direct current output
is obtained after removal of the high frequency components by a suit-
able low pass filter. If the two signals are in phase, then the output is
positive; if they are 180° out of phase, the output is negative; and if
they are of intermediate phase, then the output is reduced and is zero
when the phase difference is 90°. The intensity of the output thus
depends on both the intensity and phase of the nuclear signal input.
If the nuclear and Bl frequencies are not the same, then their relative
phase fluctuates with time and the output becomes a wave oscillating
at the difference frequency. The phase information is still retained
since the output will have some phase angle relative to Bl immedi-
ately after the 90° pulse, i.e. at the start of the decaying output. The
spectrometer output after phase-sensitive detection is thus oscillatory,
referred to Bl and with a much reduced frequency. It decays with time
characterized by T2* (Fig. 5.10). We should also note that this result
is obtained if the spectral frequencies are all higher than the B1
frequency or all lower, though the order of the spectral lines is reversed
between the two cases. Evidently, we cannot have some lower and
some higher as this would mix the resonance positions up so as to
prevent sensible analysis of the spectrum.
(a) (b)
(c) (d)
Figure 5.10 The possible range of outputs when the nuclei are exactly on
resonance with Bl and are in-phase (a) or 90° out-of-phase (b). If the nuclear
and Bl frequencies are different, then the output is at the difference frequency,
but starts on a maximum if they start in phase (c) or a minimum if they start
90° out-of-phase. Any intermediate phase is possible.
www.pdfgrip.com
The detection system 141
www.pdfgrip.com
142 The spectrometer
(a) Intensities
in memory
F
Nyquist
N, -A/, A/, -A/,
Frequency
(b)
F±F/2 A/,0,-A/,0,A/,0,-/V
-N r
Memory Location
Figure 5.11 (a) How a waveform at the Nyquist frequency, F, gives alternate
positive and negative values of the number N corresponding to peak voltage.
This assumes that the wave and the computer memory sweep are in a certain
timing relationship, (b) A waveform of lower frequency, F-F/2, gives the
same pattern as a waveform of frequency higher than the Nyquist frequency
by the same amount, F + F/2.
5.5.3.1 Folding
There are two sorts of folding, or aliasing, of signals and noise. One
is due to the ambiguity described above where the computer cannot
distinguish between signals higher or lower than the Nyquist frequency.
This means that the spectrometer has to be set up with all nuclear
www.pdfgrip.com
The detection system 143
1 acquisition
(a)
(c)
(d)
200 150 100 50 6
13
C/ppm
Figure 5.12 (a) An 100.62 MHz 13C NMR FID obtained from a single pulse
applied to carvone in CDC13. (b) The resulting Fourier transform, (c) The
resulting FID after adding together 200 FIDs. (d) The resulting Fourier trans-
form of the sum of 200 FIDs.
www.pdfgrip.com
144 The spectrometer
above. There is, however, a difference between them even if they are
equally separated from Bl in that the higher frequency component
reaches its first null beat with Bl earlier than the lower frequency
signal. In other words, they are phase shifted and it is possible to
discriminate between them using two phase-sensitive detectors (Fig.
5.13). This quadrature detector as it is called, then gives two outputs
with phase in quadrature which are collected separately in memory
and then Fourier transformed together so that all the signals appear
in their correct place relative to Bl and without phase anomalies. It
will be clear from Fig. 5.13 that this eliminates one kind of folding
and also allows us to place Bl in the centre of the spectral range. The
maximum frequency which we need now is only half that needed for
a single phase-sensitive detector for the same spectral range. This
means that the cut off frequency of the low pass filter can be halved
which in turn means that less noise is collected, halving the bandwidth
giving an increase in signal to noise ratio of \2 for an equal number
of FIDs. This is the normal method of operation.
I Reference] ^
| 90 ° Phase | low pass
I S™er I filters
»- PSD *-
(a) | Signal in
-CON
Single PSD noise Spectral range-y—[noise
(b)
,-''' 2 0 2 ^*,
i I I 1^
Quad. PSD noise]—»F Spectral range-y—[noise |
(c)
Figure 5.13 A quadrature detection system. The two boxes marked PSD are phase-sensitive detectors of
the type described in Fig. 5.9. The same reference is applied to both but one is made to be 90° out of
phase with the other. The two low frequency outputs are filtered and then converted to numbers and fed
sequentially to the computer memory. The lower part of the figure demonstrates the difference between
operating with a single or quadrature phase-sensitive detector. In the first case the spectrum range is 0 to
the Nyquist frequency -CON Noise folds in from both ends of the spectrum but is particularly troublesome
on the low frequency side because the low pass filter allows noise through without attenuation from
frequencies of 0 to +CON Hz. With the quadrature system the filter can cut off at half the frequency and
both the high and low frequency noise is attenuated. The dotted lines represent the fraction of signal with
a given frequency which pass through the filter into the receiver.
www.pdfgrip.com
The detection system 145
www.pdfgrip.com
146 The spectrometer
continues without any discontinuity for ever. This is the basis of many
NMR experiments. This means that it is possible to phase shift the
frequency. In other words, shift the sine waves by any angle, frequently
90°, 180°, or 270°. In terms of the rotating frame, the maximum can
then be at y\ -Jt', or -yf instead of jc, see Fig. 5.8.
It is not only the transmitter that has variable phases. As a direct
consequence of quadrature detection, the receiver has phase encoding
also. The quadrature detector produces two outputs, 1 and 2. These
outputs are stored at two locations in the computer, A and B. The
way that the outputs are transferred to the computer provides a mech-
anism for receiver phase cycling. This is summarized in Table 5.1. For
example a y receiver phase means that output 2 is subtracted from
location A and output 1 is added to location B.
Extensive use is made of phase changing in producing the required
changes in magnetization. We will meet many examples of this in
Chapters 7 and 8. Phase is also cycled, i.e. changed systematically over
a group of 4, 8, 16, or 32 spectra in order to minimize artefacts. Even
the routine observation of an NMR spectrum involves the use of the
CYCLOPS phase cycle to remove images due to errors in the quad-
rature detection by averaging the response from imperfectly balanced
receivers.
The CYCLOPS phase cycle is given in Table 5.2. The phase of the
transmitter pulse is cycled as is that of the receiver. The result is to
considerably reduce the size of the quadrature image.
Table 5.2 The CYCLOPS phase cycle for reducing quadrature images.
www.pdfgrip.com
Production of the spectrum 147
produces a number within the range ±(215 -1) or ±65 535 with the
sixteenth bit being used to indicate the sign. This gives a very large
dynamic range, permitting weak signals to be detected in the presence
of very strong signals. However, this can be compromised by incom-
petent spectrometer operation. The receiver gain has to be set to use
most of this digitization range, but not so high that the signal ever lies
even slightly outside the range. For example, an analogue signal of
strength +65 536 turns on the sign bit as well and produces a number
which the computer reads as -65 535. The resulting distortion of the
FID produces a wave in the baseline in mild cases, and extra signals
in more extreme cases (Fig. 5.14).
(b)
(a)
5.0 4.0 3.0 2.0 1.0 0.0
1
H/ppm
www.pdfgrip.com
148 The spectrometer
f («,( -CflQ )
Figure 5.15 A plot of the number of nuclei with angular frequency a)k different
from the resonant frequency o>0 for a small time interval t. The function nk =
f(a)k-a)^) is the line shape function.
At a small time t after the Bl pulse, those nuclei with angular velocity
a)k will have moved an angle 0^ from those nuclei exactly at resonance,
O)Q (Fig. 5.16). We can resolve the magnetic moments of the nk nuclei
along the WQ direction and at right angles to it, and, following stan-
dard alternating current theory, we shall differentiate the normal
component by the prefix i, where i = V-l. We have that
§k = ((Ok - 0)Q)t
cos 0k
sin 6k
Figure 5.16 How the total magnetic moment of the group of nuclei nk with
angular velocity a)k is displaced an angle 0^ from the on-resonance nuclei time
t after a Bl pulse has produced magnetization in the xy plane.
www.pdfgrip.com
Production of the spectrum 149
F(w) = f/(0e-io>'d/
—00
www.pdfgrip.com
150 The spectrometer
C00
Figure 5.17 The absorption and dispersion mode signals available at the
output of the phase-sensitive detector
The Fourier transform of the output data thus contains two compo-
nents, often called real (R) and imaginary (I), which correspond to
the u and v components. The inverse transform is also possible, i.e.
/(0 = T-r^He-ia"d/
^IT -oo
and this should be compared with the relations (5.1) and (5.2) derived
above.
www.pdfgrip.com
Production of the spectrum
l + r 2 2 (o> 0 -a>) 2 = 2
i.e. when T22(a)0 - a))2 = 1
or
0)0-0) = ± —-
1
2
www.pdfgrip.com
152 The spectrometer
Figure 5.18 The relationships between time and frequency domains. The
period P of the FID gives the position of the line. The rate of decay T2* gives
the linewidth and the initial amplitude A gives the line its area and therefore
its intensity proportional to
www.pdfgrip.com
Production of the spectrum 153
Figure 5.19 Equivalent time- and frequency-domain plots, the latter being
shown on the right of the figure. The lowest trace is a cosine wave of period
P, which gives a single infinitely narrow line at frequency IIP. (a) An expo-
nential decay It = /Oe~m, which gives a (m)2
Lorentzian line with sharp top but wide
skirts, (b) A Gaussian decay 7, = /Oe~ , giving a line with narrow skirts but
a thickened top. (c) A super-Gaussian decay /, = /Oe~(m)4, which introduces
flanking waves at the base of the peak. (From Akitt (1978) /. Magn. Reson.,
32, 311, with permission.)
least 10 kHz. Thus the use of short Bl pulses ensures that all the nuclei
in a sample, whatever their chemical shift, are swung around Bl by
the same angle. Intensity distortion only occurs if the chemical shift
range is very large. The same comments apply to a train of pulses,
except that the frequency coverage is now discontinuous. This has little
effect since the time between pulses will normally be longer than T2*
and the spacing between the peaks of energy will be less than a
linewidth. In all cases we will have present simultaneously in the
output, signals due to all the chemically different nuclei in a sample,
and the output will be complex and so require mathematical analysis
to obtain a spectrum.
It is also possible to produce a frequency-selective pulse. The spec-
trometer output has to be much reduced so that a 90° pulse has to be
very long. In such a case, the frequency spread of the pulse will be
only a few hertz and one can choose nuclei of a particular chemical
shift to precess around Bl without affecting any other nuclei in the
www.pdfgrip.com
154 The spectrometer
CO
ffl
CO
CO
Figure 5.20 Time- and frequency-domain equivalents of multiple-frequency
responses containing three, two, four and six components. The time-domain
signals for the lower three are relatively simple because the components have
been chosen to be of equal amplitude and regularly spaced. If the spacing
and amplitudes are irregular, the time-domain signal becomes much more
complex.
www.pdfgrip.com
Production of the spectrum 155
Figure 5.22 The ideal Fourier transform experiment, which allows the spins
to relax to equilibrium before successive pulses are applied. The data are
collected until the memory is completely traversed and all memory locations
contain data.
www.pdfgrip.com
156 The spectrometer
In practice, the data size for the FID should be chosen so that
the FID is undetectable between 40 and 80% of the way through the
dataset being collected. If the FID finishes earlier, then the noise in
the remainder of the dataset will be converted into extra noise in the
final spectrum. Also if Tl = T2, time is being wasted during the noise
collection. Alternatively, if the FID continues beyond the end of the
collected data, resolution is being lost and care is necessary if zero
filling is going to be used, as described in the next paragraph, since
this will produce a step in the data.
The quality of the spectrum is affected by the number of data
points used for both acquisition and Fourier transformation. These
do not need to be the same as it is possible to increase the data size
after the end of accumulation of the FID by adding a block of zeros.
This is called zero filling. If an inadequate number of datapoints is
used for the acquisition of the FID, then resolution is lost; compare
Fig. 5.23(a) with 5.23(d). In the case shown in Fig. 5.23(a), the FID is
non-zero at the end of the data set and attempts to improve the
resolution by adding zero filled datapoints does result in a sharpening
of the signal, but also in ringing on either side of the signal (Fig.
5.23(b)). The result of increasing the number of datapoints used for
acquisition results in the FID being fully contained in the dataset, and
the resolution gives what at first sight is an acceptable spectrum
(Fig. 5.23(c)). However, after zero-filling, the peak shape is better
mapped, and more importantly the relative heights of the two
signals is now correct (Fig. 5.23(d)). A further increase in data size
does not improve resolution, but can improve the appearance of
spectra as peaks which were previously defined by a few data points
will now be defined by more and will have a smoother appearance,
the technique being equivalent to interpolating points in the frequency
domain.
The waiting time needed to allow the spin system to come to equi-
librium is a waste of our equipment since it is quiescent while waiting.
It was quickly found that the signal-to-noise ratio can be maximized
in a given time if we adopt the procedure: pulse - acquire data - pulse
- acquire data, and omit the waiting time. Thus we allow the computer
acquisition time to determine the pulse repetition rate. This time will
usually be much shorter than the ideal of Fig. 5.22, so that some trans-
verse magnetization will still exist at the time of each pulse after the
first. In fact, the later 90° pulses will tip the magnetization through
the xy plane and the intensity will be reduced. The signal strength
is then optimized by reducing the length of the pulse, and so the
pulse angle, until a maximum steady-state output is achieved.
The pulse lengths commonly found correct for this type of data
collection are in the range of 40° to 30°, depending on nuclear isotope
www.pdfgrip.com
Rapid multiple pulsing 157
(a) (b)
(c) (d)
Figure 5.23 Examples of the effect of zero-filling. In each case only one doublet from a 162 MHz 31P AX
NMR spectrum is shown. A spectral width of 10 204 Hz was used, (a) A 16K FID transformed into a 8K
real spectrum. The angular appearance of the spectrum is due to the digitization interval being 1.3 Hz,
while the separation of the doublet is 5 Hz. (b) A 16K FID, increased to 128K by adding zeros, trans-
formed into a 64K real spectrum. The extra digitization has resulted in a reduction of the linewidth, but
the FID had not gone to zero by the end of the dataset and the result is that there is ringing on either
side of the signals, (c) A 64K FID transformed into a 32K real spectrum. This is much better, but due to
the lack of digitization, the relative heights of the two signals are wrong, (d) A 64K FID, increased to
128K by adding zeros, transformed into a 64K real spectrum. This spectrum finally gives true heights and
linewidths to the signal.
www.pdfgrip.com
158 The spectrometer
optimum pulse length will be different for each. This means that there
will be a distortion of the intensity of the signal from each type of
nucleus, which fortunately is not too grave a disadvantage in many
cases since we know a molecule is made up of integral numbers of
each type of atom. If precise quantitative data are essential, then the
pulse sequence of Fig. 5.22 must be used or a means found of reducing
Tl in our sample.
www.pdfgrip.com
Manipulation of collected data
(d).
(0.
(b)
(a)
60 50 40 30 20
13
C/ppm
Figure 5.24 The effect of the exponential multiplication process on the 100.62 MHz 13C NMR signal of
bromocyclohexane in (CD3)2CO at -16°C. (a) A transformed spectrum where no window function has
been applied, (b). A transformed spectrum which has been broadened by multiplying by a decaying
exponential with a time constant corresponding to a broadening of 1 Hz. (c) As b, but 4 Hz broadening,
(d) As b, but 10 Hz broadening. Note that as the broadening is increased, the broad signals at 8 26.5, 38,
and 53 become more clearly seen, but the coupling on the CD3 signal of the solvent at 8 20.4 becomes
more and more obscured.
www.pdfgrip.com
160 The spectrometer
www.pdfgrip.com
Manipulation of collected data 161
(d).
(0-
(b)_
(a)-
2.70 2.60 2.50 2.40 2.30
1
H/ppm
Figure 5.25 The influence of using Gaussian resolution enhancement on a *H NMR spectrum using equa-
tion (5.6). (a) No enhancement, (b) A T2 of -0.637 s, and b = 0.157. (c) A T2 of -0.318 s, and b = 0.076.
(d) A T2 of -0.59 s, and b = 0.039.
www.pdfgrip.com
162 The spectrometer
www.pdfgrip.com
Manipulation of collected data 163
and Tukey have shown that, provided all the data are contained in a
memory with number of locations equal to a power of 2 (e.g. 1024,
2048, 4096,..., etc., or IK, 2K, 4K,..., etc., locations), then the process
can be factorized in such a way as to remove many redundant multi-
plications and can be carried out in the memory containing the data,
with only a few extra locations being needed to store numbers
temporarily. The process is thus fast and economical of memory space.
The existence of the Cooley-Tukey algorithm and of minicomputers
were both necessary before FT NMR could become a commercial
possibility. In fact, the current generation of dedicated computers are
so advanced that they can perform a transform on accumulated data
in another part of memory in order to observe how the accumulation
is proceeding while at the same time more data are being added to
the FID. The sine and cosine multiplications produce, in effect, the
two components of the nuclear magnetization in phase with and
normal to Bl9 and these are kept separate in the memory. The final
result of the transform is thus two spectra, which should in principle
be the absorption and dispersion spectra. This assumes that the Bl
signal fed to the phase-sensitive detector is exactly in phase with the
nuclear signal, a condition that it is very hard to realize in practice.
Thus the two sets of spectra contain a mixture of dispersion and
absorption. In addition, we cannot be sure that we have started to
collect data immediately after the Bl pulse has ended; indeed, to avoid
breakthrough of the strong pulse signal, we may have purposely
delayed the collection of data, commonly by one dwell time. This
means that the starting point for each frequency component differs,
since each type of nuclear magnet will have precessed by a different
amount by the time data collection starts, their phases are different
and the degree of admixture of the two spectra varies across the spec-
tral width. The two components thus have to be separated, a proce-
dure known as phase correction, which is an essential part of FT
spectroscopy. The two halves of the sets of spectra are weighted and
mixed according to the formulae below.
new disp = (old disp) cos0 - (old abs) sinO
new abs = (old abs) cosO + (old disp) sinO
where 'abs' means one half of the data and 'disp' the other. The multi-
plication is carried out point by point through the data and 0 is adjusted
until one line is correctly phased, i.e. purely absorption in one half of
memory and purely dispersion in the other half. If not all the lines
are then correctly phased, it is necessary to repeat the correction but
allow 0 to vary linearly as a function of position in memory. It is best
to arrange that 0 = 0 for the correctly phased line and allow it to vary
to either side. This second process is then continued until all the lines
are correctly phased. Phase correction can be done automatically by
the computer, though the operator may intervene if needed. The
process is illustrated in Fig. 5.27.
www.pdfgrip.com
164 The spectrometer
real imaginary
before
phasing
after
phasing
Figure 5.27 Phase correction. The upper pair of spectra are the two results
of the transform operation with scrambled phase of the lines of the AB2 spec-
trum. A two-parameter phase correction gives the purely absorption spectrum
in one half and the purely dispersion spectrum in the other.
www.pdfgrip.com
Manipulation of collected data 165
5.8.3.2 Integration
The initial intensity of a FID is proportional to the number of nuclei
contributing to that signal and this transforms as the area of the
Lorentzian absorption. An integral of the spectrum (we always, of
course, refer to the absorption spectrum) then will tell us how many
nuclei contribute to a given line and can give us invaluable quantitive
data about a molecule - in the absence of relaxation effects discussed
in Chapter 4. The integration is carried out simply by adding the
numbers in successive memory locations. Where there is no resonance,
this sum will remain constant, but will increase in the region of any
resonance. The integral then forms a series of steps, rising at inter-
vals, with each rise corresponding to a resonance, see Fig. 5.28. For
the plot to be vertical and horizontal as shown, it is necessary that the
baseline is at zero intensity. Means are thus provided to correct the
baseline to give the most acceptable integral.
7.6 7.4 7.2 7.0 6.8 6.6 6.4 6.2 6.0 5.8 5.6
1
H/ppm
Figure 5.28 A 400 MHz *H NMR spectrum and its integral of 2,6-
dichlorophenol in CDC13. The ratio of the heights of the steps is approxi-
mately 2 : 1 : 1 as required for H3,H5:H4:OH.
www.pdfgrip.com
166 The spectrometer
5.8.3.3 Expansion
The typical monitor screen is rather small to be able to observe much
detail, and means are provided to enable a part of the spectrum to be
selected and expanded to fill the screen to assist in the various oper-
ations needed to improve a spectrum and, for instance, show up small
splittings.
5.9 QUESTIONS
www.pdfgrip.com
Making the spins dance ®
We are now going to discuss how we can predetermine the behaviour
of the nuclear magnetization using pulses of various lengths and phase
and carry out a variety of NMR experiments.
6.1 DECOUPLING
www.pdfgrip.com
168 Making the spins dance
1 I I I I I
3.,30 3.25 3.30 3 .25 3.,30 3.25
1H/ppm 1H/ppm 1H/ppm
Figure 6.1 The 1H NMR spectrum31 of P(OCH3)3. (a) Normal, with 31P
coupling, (b) With irradiation at the 31 P frequency to completely decouple the
phosphorus, (c) With low-powered P decoupling.
www.pdfgrip.com
Decoupling 169
(0-
(b)
(a):
7.6 7.4 7.2 7.0
16.8 6.6 6.4 6.2 6.0 5.8 5.6
Figure 6.2 An example of homonuclear double irradiation, (a) The normal 400 MHz !H NMR spectrum
of 2,6-dichlorophenol in CDC13. (b) As 4a, but with B2 placed on the doublet due to H3 and H5 at 8 7.24.
Note that the triplet at 8 6.80 due to H has been reduced to a slightly broadened singlet, (c) As a, but
with B2 placed on the triplet due to H4 at 8 6.80. Note that the doublet at 8 7.24 due to H3 and H5 has
been reduced to a slightly broadened singlet. There is also a decoupling spike at 8 6.80 due to leakage of
the decoupling frequency into the receiver. The OH proton at 8 5.89 is not affected.
www.pdfgrip.com
170 Making the spins dance
Decoupler On
RXon
.A/D
one dwell time interval i -Time
Figure 6.3 Time-shared homonuclear decoupling. Three parts of the spec-
trometer have to be switched on and off independently though in a particu-
lar order. The receiver alone is switched on for sufficient time to establish an
output. This output is fed to the analogue-to-digital (A/D) converter, which
starts to convert the voltage into a number at a precise time, set by the dwell
time in use, and takes a finite time to make the conversion, 5 to 20 JJLS. The
decoupler is switched on during the A/D conversion process and remains on
until a little while before the receiver is switched on again. The effect of the
decoupler may be modified by reducing the length of each transmitter pulse
(decoupler on) as required.
www.pdfgrip.com
Decoupling 171
(b)
(a).
8.50 8.25 8.00 7.75 7.50 7.25 7.00 6.75 6.50 6.25
1
H/ppm
www.pdfgrip.com
172 Making the spins dance
(C)-
31
Figure 6.5 The 162 MHz :
P spectrum of triethyl phosphite, (CH3CH2O)3P.
(a) Obtained without H irradiation, showing coupling to the methylene
protons broadened by the coupling to the methyl protons, (b) The methyl
protons irradiated, which gives a well resolved septet due to coupling to the
methylene protons, (c) The methylene protons irradiated, which selectively
removes their influence from the spectrum.
www.pdfgrip.com
Decoupling 173
(c)-
(b)-
(a)
i
200 100 0
8/ppm
Figure 6.6 The 100.62 MHz 13C NMR spectrum of carvone in CDC13.
(a) Obtained with WALTZ decoupling,
13
(b) Obtained with the !H decoupling
frequency placed at 8 10. The C nuclei attached to high frequency protons
near 810 show big reductions in 1J(13C1H). (c) Obtained without *H
decoupling.
www.pdfgrip.com
174 Making the spins dance
6.1.1.4 WALTZ
In order to have some appreciation of how composite pulse decou-
pling works, let us examine the effect of applying a sequence of 180°
pulses to X H while observing 13C of a compound such as CHC13. In
the absence of !H pulses or decoupling, the 13C NMR signal will consist
of a doublet, due to coupling to 1H. The doublet arises because half
the 13C nuclei are attached to 1H nuclei with mI = 1/2 and the other
half are attached to 1H nuclei with ml = -1/2. If we now apply a 180°
pulse to XH, then we reverse the ml values of the !H nuclei. The 13C
nuclei which were attached to *H nuclei with ml = 1/2 are now attached
to !H nuclei with ml = -1/2 and vice versa. If we were to apply a
sequence of 180° pulses, each 13C nucleus would see a !H nucleus
which was rapidly reversing its spin. The result is XH decoupling.
The experiment described above is inefficient and composite pulses
are preferred, see section 6.2. Generally, the WALTZ-16 pulse
sequence is used. It is built up of composite pulses based on the
sequence (90°)JC(180°)_JC(270°);c. This is given the shorthand notation
1 "2T 3, with 4 being used for a (360),, pulse. The numbers represent
multiples of 90° and the bar over the 2 represents a reversal of phase
to produce a 180° phase shift. The shorthand notation leads to the
name WALTZ as the dance is based on a 1-2-3 step sequence. There
are several variations depending on the length of the sequence used,
but it is generally the WALTZ-16 sequence which is preferred.
www.pdfgrip.com
Decoupling 175
6.1.1.5 GARP
In order to increase the bandwidth of the decoupling, it is necessary
to use flip angles which are not multiples of 90°. This has the effect
of broadening the bandwidth, but makes the decoupling efficiency less.
The bandwidth doubles, but the residual splitting increases to around
0.3 Hz. The GARP-1 pulse sequence is based on the pulse sequence
RR R 7F, where in terms of pulse angles
R= 30.5 352 257.8 2683 69.3 622 85.0 ~9L8 134.5 25O
66.4 ~45^ 25.5 72J 119.5 1382 258.4 "649 70.9 772
98.2 133.6 255.9 65.5 53.4
Typically the X decoupler is attenuated so that the 90° pulse is
around 70 JJLS. The acronym GARP comes from Globally optimized
Alternating-phase Rectangular Pulses.
www.pdfgrip.com
176 Making the spins dance
phase with that after the shorter pulse. If we arrange our computer
to give us a positive-going absorption peak from the FID following
a 90° pulse, then after a 270° pulse we will get a negative peak. If,
instead, we use a 180° pulse, we create no Mxy but turn the excess
low-energy spins into the high-energy state, -Mr They will relax to
their normal state with characteristic time Tl and the magnetization
will change from - Mz through zero to + Mz (Fig. 5.35). This, of course,
produces no detectable effects. However, if, at some time r (do not
confuse this with rc) after the 180° pulse, we apply a second pulse of
90°, we will create magnetization Mxy equal in magnitude to Mz at that
instant. The spectrum that results from processing the FID will be
negative-going if Mz is still negative (as if part of the magnetization
had undergone a 270° pulse) and positive-going if Mz has passed
through zero (Fig. 6.7). A series of spectra are obtained in this way
for a number of different values of r and the intensity of the resulting
peaks plotted as a function of r, so allowing us to extract a value for
7\. The equations governing the behaviour of the transverse and longi-
tudinal magnetization Mxy and Mz and their return to equilibrium
following a 180° Bl pulse are
(M,X = (M,)Jl-2exp(^)|
L \M/J
FT
FT
Figure 6.7 (a) Production of a nuclear response using a (90°),, pulse. The FID and its transform are repre-
sented on the right of the figure, (b) A similar response is obtained after a (270°),, pulse, but the FID is
180° out of phase, giving an inverted transform, (c) A 180° pulse gives no xy magnetization but places Mz
in a non-equilibrium position opposing the field. This magnetization relaxes back to its equilibrium value
and no transverse magnetization is produced at any time throughout this process.
www.pdfgrip.com
Decoupling 177
Figure 6.8 How the pulse sequence 180°-wait T-90° can be used to perturb Mz and follow what then
happens to Mr
www.pdfgrip.com
178 Making the spins dance
Figure 6.9 The full population inversion experiment. A series of spectra are
obtained at different T. A plot of their intensities as a function of time gives
the rate of relaxation, from which Tl can be derived. (After Martin et al
(1980) Practical NMR Spectroscopy, John Wiley and Sons Ltd, reprinted with
permission.)
6.12.2 Measurement of T2
The contribution of magnetic field inhomogeneity to T2 will usually
be of the order of 3.0 to 0.3 s for a high-resolution magnet. Thus,
provided T2 is less than about 0.003 s, it can be measured directly
either from the line width or from the rate of decay of the FID. In the
latter case we have to suppose that there is only a single resonance.
Accurate measurement of longer T2 is made using a spin-echo exper-
iment. This depends upon the rate at which the FID decays being
faster than the rate of longitudinal relaxation R{. The decay is due to
the T2* mechanism, which has two components. The loss of phase
coherence between the spins in the xy plane is caused both by the
random relaxation field and by the fact that the homogeneity of the
magnetic field within the sample is not perfect, so that nuclei in
different parts of the sample have different precession frequencies.
There is an important difference between these two relaxation
processes: the first is entirely random, and so unpredictable; whereas
www.pdfgrip.com
Decoupling 179
the second acts continuously and is constant at each part of the sample,
so that we can in principle correct for this relaxation contribution. We
assume that the random contribution to relaxation is negligible. If we
apply a 90° pulse, all the magnetization Mxy is in the xy plane, but
spins in different parts of the sample have different angular velocities
due to the inhomogeneities in the magnetic field, and so some move
ahead of the average and some lag behind. We wait a period of r
seconds for the spin distribution to evolve; the value of Mxy will
decrease, and may even become zero if the field inhomogeneity is
large. We then apply a 180° pulse, and all spins precess around Bl to
the other side of the xy plane. They do not take up their mirror posi-
tions. We have thus put the faster spins at the rear and the slower
spins in front. They continue to precess, but after further time r
seconds they come into phase again and the magnetization Mxy is again
a maximum. The 180° pulse is a refocusing pulse and the intensity of
the spectrometer output rises following this pulse to a maximum r
seconds after and then decays again. Another 180° pulse will refocus
the magnetization, which indeed can be refocused indefinitely, given
perfect pulses. The refocusing does not work for the random relax-
ation process, and so the duration of the experiment is limited by the
intrinsic transverse relaxation. Indeed, the train of echoes decays in
intensity at a rate determined by the real T29 which can be measured
from a plot of intensity versus time. The experiment is summarized in
Fig. 6.10. Each echo is effectively two back-to-back FIDs. It is possible
to split them and Fourier transform each so that, if the sample contains
several resonances, then T2 can be measured separately for each from
the resulting absorption spectra. If spin-spin coupling is present,
however, then this does not work and modulation of the echoes is
produced, which, as we shall see, proves useful in multidimensional
spectroscopy.
www.pdfgrip.com
180 Making the spins dance
(a)
(b)
Figure 6.10 Illustrating the Carr-Purcell pulse sequence for measuring T2. The behaviour of the spins is
shown relative to the rotating frame, as if they were stationary. The (90°)^. pulse produces magnetization
Mxy, which then decreases as the spins move apart, s = slow, f = fast. The (180°)^ pulse alters the relative
positions of the slower and faster spins, which now close up again and Mxy increases, reaching a maximum.
The spins are said to be refocused. The output then decreases again. Fig. 6.10(b) shows diagrammatically
how the first and second (the echo) FIDs appear.
www.pdfgrip.com
Composite pulses 181
www.pdfgrip.com
182 Making the spins dance
Figure 6.12 The path followed by the tip of the magnetization vector during
a (800)JC(160°)y(800).c pulse sequence is shown by the solid line. The pulse
sequence has converted the z component of Mz to -0.998MZ compared with
-0.940M., produced by a single 160° pulse.
180° pulse corresponds to a 160° pulse. The first pulse rotates the
magnetization vector 80° around the x axis towards the xy plane. The
second pulse rotates the magnetization vector 160° around the y axis.
The third pulse rotates the magnetization vector 80° around the x axis
and places it close to where it would have been if a true 180° pulse
had been applied.
In practice, it has been found that a slightly different sequence, the
(9Q0)x(24Q°)y(9Q°)x composite pulse sequence works best.
Pulse sequences are frequently used with delays between pulses. This
can cause very bad phasing problems. Consider what happens if a
www.pdfgrip.com
Refocusing pulse 183
Figure 6.13 The path followed by the tip of the magnetization vector during
a (800)^(800)y pulse sequence is shown by the solid line.
www.pdfgrip.com
184 Making the spins dance
due to this will also be refocused (Fig. 6.15). Comparison of this pulse
sequence with that used for T2 determination shows that it is the
Carr-Purcell pulse sequence.
Figure 6.15 The effect of the (180°) refocusing pulse on a collection of nuclei, a, b, c, d, and e, with
different resonance frequencies, (a) The alignment of the nuclear spins immediately following a (90°)^
pulse, (b) After a time T, the nuclei have fanned out with nuclei a and b rotating faster than the rotating
frame, and nuclei d and e rotating slower. Nuclei c are on resonance, (c) The effect of the (180°)v refo-
cusing pulse, (d) After a time T, the nuclei have refocused.
www.pdfgrip.com
Refocusing pulse 185
p1 d1 p1
(a) (b)
Figure 6.16 (a) The DANTE pulse sequence, (b) The frequency distribution generated by the DANTE
pulse sequence.
a perfect square pulse. There is a rise and fall time. The 180° DANTE
pulse has to be calibrated, and each component can be significantly
longer than 180°/n when the required length is less than 1 JJLS.
(c)
Figure 6.17 (a) A rectangular pulse and its Fourier transform, (b) A sin xlx
pulse and its Fourier transform, (c) A Gaussian pulse and its Fourier trans-
form.
www.pdfgrip.com
186 Making the spins dance
-x
1
H
d1 p1 d2 p2 d2 p3 d3 P5
x
13C
p4
Figure 6.18 The BIRD pulse sequence, dl 1is a relaxation delay, 57\ being ideal. d2 = 1/27 s. d3 is a delay
for the -z magnetization generated in the H nuclei attached to 12C to decay to zero, pi, p3, and p5 are
90° pulses, while p2 and p4 are 180° pulses. The phases of the pulses are given above them. The actual
BIRD pulse sequence comprises pi, d2, p2, p3, and p4. d3 and p5 have been added here to demonstrate
the consequences of applying the pulse sequence.
www.pdfgrip.com
Refocusing pulse 187
(a)
180°(13C)
(b)
Figure 6.19 Part of the BIRD pulse sequence to show the behaviour of the nuclear spins in the xy plane,
(a) The behaviour of the 1H13C doublet, (b) The behaviour of the 1H12C singlet.
www.pdfgrip.com
188 Making the spins dance
using phase cycling of the type described above, with the difference
that phase cycling requires many acquisitions to complete a cycle
whereas field gradient pulses do the job in one acquisition. They are
used particularly in biochemical spectroscopy.
6.4 QUESTIONS
www.pdfgrip.com
NMR spectra of
exchanging and
reacting systems
7
7.1 SYSTEMS AT EQUILIBRIUM
One of the most important contributions that NMR has made to chem-
istry is the insight it has given into the dynamic, time-dependent nature
of many systems, particularly those which are at equilibrium or where
simply intramolecular motion is involved. Spectroscopy based on
higher-frequency radiation, such as classical infrared (IR) or ultra-
violet (UV) Spectroscopy, has given mostly a static picture because the
timescale of many processes is slow relative to the frequency used.
However, the lower frequencies used for NMR and the smaller line
separations involved, coupled with the small natural linewidths
obtained, means that many time-dependent processes affect the spectra
profoundly. As an example, we consider the spectroscopic behaviour
of ethanol in Fig. 7.1. The proton spectrum of 50% ethanol,
HOCH2CH3, in CDC13 is a methyl triplet due to coupling to the CH2,
an OH triplet for the same reason and a doublet of quartets for the
methylene protons. Any acidic impurity catalyses interchange of OH
protons between molecules:
H+
EtOH + EtOH* ^ ^ EtOH* + EtOH
The exchange of the protons results in a short break in the CH2OH
coupling path and, since the total spin of the CH2 protons in the two
molecules between which the proton jumps may not be identical, then
some of the OH protons will suffer an abrupt change in frequency.
The result is to introduce uncertainty into their nuclear frequency and
thus line broadening. In the spectrum of Fig. 7.1(b), which is of 50%
ethanol, coupling of the OH and CH2 proton signals is evident.
Addition of acid causes acceleration of the rate of proton exchange
to the extent that the OH-methylene coupling is completely lost and
only an average frequency can be detected (Fig. 7.1(c)). The lines are
now sharp. This is called the fast exchange region. When the lines
are broadened (Fig. 7.1(a)) it is called the region of intermediate
exchange rates and where coupling is fully developed is called the slow
exchange region.
www.pdfgrip.com
190 NMR spectra of exchanging and reacting systems
CH3
CH2 OH
(a)
CH3
OH
OH CH2
(c)
5 4 3 2 1 0
1
H/ppm
Figure 7.1 The 400 MHz ]H NMR spectra of ethanol. (a) 2% EtOH in CDC13. (b) 50% EtOH in CDC13.
(c) As (b), but with a drop of acid added.
www.pdfgrip.com
Systems at equilibrium 191
methyl protons are close to the OH group and one is pointing away;
see the Newman projection in Fig. 7.2. These differences are not
observed in the spectrum, however, because each proton has an
average chemical shift due to the rapid internal rotation. In certain
molecules, such conformational changes can be quite slow and then
the effects of the motion can be detected in the NMR spectra.
Following the Karplus relationship (Fig. 3.1), the interproton coupling
constants will also be different in an ethyl group; these are also aver- Figure 7.2 A Newman
aged by the rotation to the value of the average of the Karplus curve. projection about the C-C
bond of ethanol, showing
the two environments for
7.1.1 The effects of exchange on the lineshape of NMR spectra the methyl protons.
We have described exchange by such words as 'fast' and 'slow'. It is
now necessary to determine the timescale within which we can apply
these terms correctly. A set of theoretical spectra for a two-site
exchange with equal population in the two sites and no spin-spin
coupling is shown in Fig. 7.3.
We start from the situation where the nuclei spend a long time in
a given location in the molecule, Fig. 7.3(a), as shown in section 4.1,
the linewidth is given by
^1/2 = —
vT2
When chemical exchange occurs, the nuclei change their chemical
shift as a consequence, and spend a time, TC, in a given location in the
molecule. There is an extra contribution to the linewidth due to
exchange and the linewidth of the exchange broadened line at half
height, (W1/2)ex, becomes
W/2)ex = J- + J-
TtT2 TTTex
k=±
T
ex
Hence
W/2L = ^ + T
and k can be derived from the linewidth of the broadened line, (W1/2)ex,
by the simple calculation
www.pdfgrip.com
192 NMR spectra of exchanging and reacting systems
(a),
(b).
(c).
(d)-
(e).
(f)-
(g)-
(h).
(0-
(j)-
Figure 7.3 Calculated spectra for exchange between two equally populated
sites separated by 40 Hz.1 T2 values of both sites are 1 s. 1(a) k = 0.1 s~l. (b)
k = 5s- . (c) k = 10s- . (d) k = 20s- . (e) k = 40s' . (f) k = 88.8s-1.
1 1
(g) k = 200 s-1. (h) k = 400 s'1. (i) k = 800 s~l. (j) k = 10 000 s'1, where k is
the rate of exchange between the two sites, and is normally varied by varying
the temperature.
www.pdfgrip.com
Systems at equilibrium 193
As the rate of exchange increases, the lines broaden, and the trough
between the lines gradually fills. Once the lines are no longer well
resolved, e.g. Fig. 7.3(d), the use of equation (7.1) becomes subject to
error. The temperature where k is such that the trough between the
two lines rises to be level with the peaks, Fig. 7.3(e), is known as
the 'coalescence temperature'. At the coalescence temperature
TT|VA-VB|
k- j- (7.2)
where VA and VB are the frequencies of the two sites in hertz at the
coalescence temperature. This equation only applies to exchange
between two equally populated uncoupled sites and (Wl/2)o has to be
negligible when compared with I VA - VB I. In principle, very accurate
rates can be obtained at the coalescence temperature, provided
that (VA - VB) is known. As chemical shifts are temperature dependent,
(V A -V B ) must be determined by extrapolation from lower tempera-
tures, or, better, by complete lineshape fitting. Caution is necessary
when extrapolating from low temperature, as when the signals begin
to overlap and the trough between them fills, the maxima move
together (Fig. 7.4).
It must be remembered that there is not a single coalescence temper-
ature for a compound. (VA - VB) depends on the magnetic field strength,
the pair of nuclei exchanging, and the solvent. It is therefore mean-
ingless to quote the coalescence temperature for a compound.
Above coalescence, the signals of the two lines are averaged, initially
to a non-Lorentzian lineshape, but as the rate of exchange increases
further, a Lorentzian lineshape is achieved (Fig. 7.3(g)-(j)). Then the
equation
k= ^(VA ~ ^B)2 (73)
2{(W1/2)ex - (W1/2)0)
applies. This equation only applies to exchange between two equally
populated uncoupled sites. The equation must be used with caution,
as it is frequently difficult to obtain a reliable estimate of (VA - VB) at
the experimental temperature. The dependence of equation (7.3) on
(VA - vB)2 means that the line broadening of signals above the coales-
cence temperature can differ markedly for each pair of exchanging
signals.
Frequently exchange processes affect several signals from a
compound. This is illustrated here for W-methylaniline (Fig. 7.5). At
-133°C, separate 13C NMR signals are observed from all the six
aromatic carbon atoms, showing that there is slow rotation about the
Ph-N bond. At -126°C, the rotation produces a noticeable broadening
of the signals due to C2, C3, C5, and C6 as can be observed by comparing
the heights of these signals with the height of the signal due to C4. On
warming to -115°C, the signals due to C3 and C5 have averaged to
give one signal, while one broadened signal is observed for C2 and C6.
www.pdfgrip.com
194 NMR spectra of exchanging and reacting systems
62°C
40°C
36°C
32°C
26°C
10°C
-2°C
-27°C
6.0 5.5 5.0
1
H/ppm
www.pdfgrip.com
Systems at equilibrium 195
3,5
2,6
-75°C
-115°C
-126°C J Jl
-133°C
Figure 7.5 The variable temperature 25.16 MHz ^C^H} NMR spectrum of
W-methylaniline in Me2O. (Reproduced from Lunazzi et al (1979) Tetrahedron
Letts., 3031, copyright (1979), with permission from Elsevier Science.)
than the stronger one from site B. This is a direct consequence of the
law of microscopic reversibility. This predicts that
^ = * B or *A = *B£*
PE PA PA
where A:A and kE are the rates of leaving sites A and B, respectively,
and pA and pE are the populations of sites A and B, respectively.
Hence if pA = 3/?B, then according to equation (7.1), the line broad-
ening is proportional to k, the signal of B will broaden twice as much
as the signal due to A. This is apparent in Fig. 7.6(c).
Above coalescence an average signal is observed at the weighted
average chemical shift of the two signals, vav, where
_ VAPA + VBPB
Vav
PA+PB
www.pdfgrip.com
196 NMR spectra of exchanging and reacting systems
(a)
(b)
(c)
(d)
(e)
(f)
(9)
(h)
(i)
(j)
Figure 7.6 Calculated spectra for exchange between two sites with popula-
tions in the ratio 1: 3 separated by 40 Hz. T2 values of both sites are 1 s. (a)
k = 0.11 s-1. (b) k = 5 s- 1
. (c) k = 10 s'11. (d) k = 20 s-11. (e) k = 40 s'1. (f) k=
88.8 s- . (g) k = 200 s- . (h) k = 400 s' . (i) k = 800 s' . (j) k = 10 000 s'1.
1
equatorial axial
bromide bromide
The variable temperature 13C NMR spectra are shown in Fig. 7.7.
The determination of the exchange rates and ratio of isomers where
separate signals are observed is easy and accurate up to about -44°C.
At higher temperatures, where only average signals are observed, the
determination of accurate values of (VA - VB) and ratio of concentration
www.pdfgrip.com
Systems at equilibrium 197
2e 3e
6e 5e
1e 2a 4e 3a
6a 5a
1a 4a
-94 °C_ J_
-60 °C-
-44 °C-
-30 °C.
-16 °C.
-2°C-
11 °C. Al i
20 °C
55 50 45 40 35 30 25 20
13
C/ppm
Figure 7.7 The variable temperature 100.62 MHz 13C NMR spectrum of
bromocyclohexane in CD3C6D5. The 3a and 5a resonance is partially obscured
at -44°C and below by the CD3 solvent resonance. The a and e refer to the
position of the bromine atom, axial or equatorial.
www.pdfgrip.com
198 NMR spectra of exchanging and reacting systems
p1 p2 p1 p2
2
IP3 1
IP3 2
IP3 5
IP1
P -R< 4A ^=^P -R< 4A P -R< 4, ^=^P -RI< 4.
l>
p5
l>
p5
l>
po
l>
p3
K
(a) (b)
www.pdfgrip.com
Systems at equilibrium 199
Berry
pseudo-rotation
calculated
Observed
-114 °C
Pairwise
exchange
calculated
Figure 7.9 Observed and calculated 36.43 MHz 31P NMR spectra of
[Rh{P(OMe)3}5]+ in CHC1F2/CH2C12, 9:1 at -114°C, with the effect of the
protons removed by double irradiation. Arrows are used to draw attention
to the regions of the spectrum where differences are most marked.
(Reproduced with permission from Meakin and lesson (1973) /. Am. Chem.
Soc., 95, 7272, copyright (1973) American Chemical Society.)
www.pdfgrip.com
200 NMR spectra of exchanging and reacting systems
or
1 M 7 B (0)-M 7 B H
kB
-T» M/H ^
If kE » I/TV, then M,BH -0. If kE « 1/7\B, then M,B(o>) ~MZB(0).
Hence if kE » IIT^, magnetization transfer causes the signal due to
B to vanish as well as the signal due to A. If kB «1/7\B, the signal
due to B is unaffected by magnetization transfer by exchange. If
1/5Tj 8 «k E «5/7"1B, then kB can be determined by measuring
MzB(oo), MZB(0), and T*. This works well if 7\A ~ Tf. Unfortunately
if 7\A * 7\B, it is very difficult to measure T^ and T^ as exchange
occurs during the waiting period of the IT - T - ir/2 pulse sequence,
resulting in a partial averaging of 7\A and T-f.
Considerable use is made of magnetization transfer measurements
by decoupling in order to determine which signals are exchanging,
and it can be seen accidentally during NOE measurements, see section
8.2. An example of its use is to determine how the osmium atom
moves around the cyclooctatetraene ring in [Os(iri6-C8H8)(Ti4-C8H12)]
(Fig. 7.10).
Figure 7.10(a) shows the normal 1H NMR spectrum of the cyclooc-
tatetraene protons of [Os(in6-C8H8)(V-C8H12)]. The spectra in Fig.
7.10(b) and (c) are presented as difference spectra. A reference spec-
trum was also taken where the decoupling frequency was placed well
away from any signal. This was then subtracted from the spectra where
the decoupler was placed on H2 in Fig. 7.10(b) and H4 in Fig. 7.10(c).
www.pdfgrip.com
Systems at equilibrium 201
's(cod)
5.0 4.0
1
H/ppm
www.pdfgrip.com
202 NMR spectra of exchanging and reacting systems
Os ^ Os
(r|4-C8H12) (i!4-C8H12)
^.
ai
£ MM+ t W<> +L P1
^} (7.6)
/=!(/*;) y=i('>y) ^i *
www.pdfgrip.com
Systems at equilibrium 203
O—ZrH O—2
Mb:
/
H H delay time
0.002s
number of times and 7\ is averaged over both sites. This makes the
analysis of the data easier, but it also removes a possible extra compli-
cation. There is another mechanism by which one proton can perturb
the intensity of another, namely the nuclear Overhauser effect, see
section 8.2. This effect operates on the same time scale as Tl and by
choosing a temperature where k is fast compared with Rl9 it can be
neglected.
www.pdfgrip.com
204 NMR spectra of exchanging and reacting systems
around the sample holder and passing over the sample cold nitrogen,
boiled-off from liquid nitrogen, or air which have been adjusted to the
required temperature by using a controlled heater. The heater is regu-
lated by a thermocouple placed just below the sample tube. The
method is prone to temperature gradients and so inaccuracies, which
are minimized by calibrating the temperatures. This is done in two
major ways. In the first method a capillary tube, containing MeOH
for below room temperature and HOCH2CH2OH for above room
temperature, is placed inside the NMR tube. The separation between
the OH and CH3 or CH2 groups, Av ppm, is measured. The temper-
ature is related to Av by the equations
for methanol
r°C - 403.0-29.46Av - 23.8Av2
for ethylene glycol
T°C = 466.0 -101.64Av
In the second method, the NMR tube is replaced by one containing
a thermocouple in the same solvent.
www.pdfgrip.com
Systems at equilibrium 205
2.0
13,
.1.5
O)
o
1.0
www.pdfgrip.com
206 NMR spectra of exchanging and reacting systems
u
ln(£)
data from
-2- lineshape analysis
-4-
-6-
-8-
data from
magnetization transfer
-10
0.0022 0.0024 0.0026 0.0028 0.003 0.0032
Figure 7.14 The Eyring plot as applied to exchange data for Me2NCHO in
DMSO/d6-DMSO. (From Mann et al (1977) /. Magn. Reson., 25, 91, reprinted
with permission.)
www.pdfgrip.com
Systems at equilibrium 207
Figure 7.15 The 300 MHz *H NMR spectra of the vinyl diamides shown at a series of temperatures. The
solvent is d8-toluene. (From Szalontai et al. (1989) Magn. Reson. Chem., 27, 216-22, copyright (1989) John
Wiley and Sons Ltd, reprinted with permission.)
that is rotating faster. The cis isomer shows much more complex behav-
iour, though at the higher temperatures there are two corresponding
coalescence points at 355K and 333K. The quartets, however, split
further at lower temperatures, and each resonance is transformed into
a doublet of quartets due to the introduction of extra spin-spin
coupling between geminal protons on the same carbon atom. The
methylene protons have thus been rendered non-equivalent by some
further restriction in the motion of the molecule. The activation para-
meters were obtained by calculating T at the coalescence points.
To do this accurately, it is of course necessary to know the chemical
shift between the two coalescing signals. This may vary with temper-
ature, and for the present example it was found necessary to measure
the chemical shifts in the slow exchange limit over a wide range of
temperatures to ensure that the correct frequency separation had
been used.
www.pdfgrip.com
208 NMR spectra of exchanging and reacting systems
7.1.5.2 [Co4(CO)12J
In the case of systems exhibiting fast exchange, it may simply be neces-
sary to reduce the temperature sufficiently to slow down the exchange
process and cause the slow exchange limit spectrum to appear, since
this will be much more informative than the fast exchange spectrum,
which may be just a singlet. A typical example is the 17O spectrum of
the cobalt carbonyl compound, [Co4(CO)12]. In this case, the 17O NMR
spectrum was recorded rather than the 13C NMR spectrum due to
extreme broadening of the 13C NMR signals by scalar relaxation by
59
Co, see section 4.6. The carbonyl groups move around the cobalt
cluster so that they all experience the full range of chemical environ-
ments and shifts in the molecule. Spectra at several temperatures are
shown in Fig. 7.16, and it will be evident that the ambient-tempera-
ture trace is not very informative but that at -25°C all four types of
carbonyl group can be distinguished. Note that the 59Co spectrum of
this compound contains two resonances in the intensity ratio 3 :1 so
that the carbonyl exchange takes place on a cluster in which the posi-
tions and bonding of the cobalt atoms are invariant. The nucleus 59Co
is quadrupolar and the asymmetry of the environment around the
metal atoms means that the relaxation time is very short. Linewidths
are of the order of 7500 Hz but 59Co chemical shifts are very large and
so the resonances are resolved.
www.pdfgrip.com
Systems at equilibrium 209
+85 °C
+57 °C
+21 °C
-25 °C
-55 °C
O
Figure 7.16 The structure, (a), and the 54.25 MHz 17O NMR spectrum, (b), of [Co4(CO)12] in CDC13. The
carbonyl groups exchange positions, but this can be slowed down sufficiently by cooling to enable the four
types of carbonyl group to be seen. The reference is water. (From Aime et al. (1981) /. Am. Chem. Soc.,
103, 5920, copyright (1981) American Chemical Society, reprinted with permission.)
www.pdfgrip.com
210 NMR spectra of exchanging and reacting systems
www.pdfgrip.com
Systems at equilibrium 211
25°
-84°
-87° - 0.00048s
-89° 0.00105s
-93° 0.0022s
-95° 0.00275 s
-100° 0.0032s
0.0070 s
-107°
0.0115s
-108°
0.015s
-109°
0.022 s
-115°
0.12s
-120
0.20s
-128°
0.25s
-147 0 100 200
Hz
0 100 200
Hz
as it only exchanges 1 and 4 and 2 and 3 and will give rise to two
signals at the high temperature limit, rather than the observed singlet.
This left the 1,3- and random shift mechanisms, but line shape analysis
cannot differentiate between these mechanisms as both mechanisms
lead to equal broadening of the lines. This led to a disagreement
between Cotton who favoured the 1,3-shift mechanism and Whitesides
who favoured the random shift mechanism.
The argument was resolved by using saturation transfer (Fig. 7.19).
The most informative spectrum is (b) where irradiation at site 4
produces a greater reduction in the intensity at sites 2 and 3 rather
www.pdfgrip.com
212 NMR spectra of exchanging and reacting systems
www.pdfgrip.com
Systems at equilibrium 213
(c)
(b)
21
(a)
www.pdfgrip.com
214 NMR spectra of exchanging and reacting systems
(a)
(b)
(c)
Figure 7.20 The two mechanisms, (a) and (b), of exchange identified for the
fluxionality of the carbonyls and interchange of isomers of [Ir4(CO)n(PEt3)].
Letters are used to denote the carbonyls and PEt3 is abbreviated to P. (c).
The lettering used to identify the carbonyls in the minor isomer in Fig. 7.21.
+ H+ -H+ +H + -H+
www.pdfgrip.com
Systems at equilibrium 215
(C)
(a)
13
210 200 190 180 170 160 C/ppm
Figure 7.21 The effect of applying a selective 180° pulse in the pulse sequence,
relaxation time-'TT(selective) - T - ir/2, to the bridging carbonyl in the major
isomer of [Ir4(CO)n(PEt3)] in CD2C12. (a) The 100.62 MHz 13C NMR spec-
trum of the carbonyl region of the spectrum acquired with T > 57\. (b) The
100.62MHz 13C NMR spectrum of the carbonyl region of the spectrum
acquired with T = 3 JJLS. (c) The difference 100.62 MHz 13C NMR spectrum of
the carbonyl region of the spectrum after subtracting spectrum (b) from a
spectrum with T = 0.01 s. (Note that this is at higher gain.) (Reproduced with
permission from Mann et al (1989) /. Chem. Soc., Dalton., 889.)
www.pdfgrip.com
216 NMR spectra of exchanging and reacting systems
HoO
[AI(OH2)n]3+
Figure 7.22 The 60 MHz *H NMR spectrum of 3 mol H A1C13 at -47°C The
water complexed by the cation is seen 4.2 ppm to high frequency of free, bulk
water. (After Schuster and Fratiello (1976) /. Chem. Phys., 47, 1554, with
permission.)
www.pdfgrip.com
Systems at equilibrium 217
www.pdfgrip.com
218 NMR spectra of exchanging and reacting systems
10.0
9.0
tf 8.0
^
7.0
6.0
2.5 3.0 3.5 4.0
(a)
10.0
5.0
£
.0^
^ 0.0
-5.0
www.pdfgrip.com
Systems at equilibrium 219
www.pdfgrip.com
220 NMR spectra of exchanging and reacting systems
experimental calculated
o^ °r*
£.0 u 10,000s
-7°C 70s 1
-9°C 48s 1
-16°C 13s1
-19°C 7s-1
pp or*
-OO U OS'1
I I I T I I I
20 0 -20 3J 0 -3J
/Hz
Figure 7.25 The 22.6 MHz 13C NMR spectrum of [(CH3)3C6Li]4 with the effect
of the protons removed by double irradiation at different temperatures in
cyclopentane solvent. Only the resonances of the a-carbon atom are shown.
(From Thomas et al (1986) Organometallics, 5, 1851, copyright (1986)
American Chemical Society, reprinted with permission.)
www.pdfgrip.com
Systems at equilibrium 221
www.pdfgrip.com
222 NMR spectra of exchanging and reacting systems
100 Hz
M W
www.pdfgrip.com
Systems at equilibrium 223
Ay-J»r(«ln£) T
www.pdfgrip.com
224 NMR spectra of exchanging and reacting systems
(Me2N)2CO
Observed Calculated Observed Calculated
P/MPa /c/s 1 P/MPa
202 359.0
106 323.5
298.3
Figure 7.27 The 200 MHz 1H spectra of acetonitrile-d6 solutions containing Be(II) and (NMe2)2CO (left)
or Me2SO (right), obtained as a function of pressure. The actual spectra are on the left and the calculated
envelopes are on the right. (From Merbach (1987) Pure Appl Chem., 59, 161, with permission.)
These values are also given in Table 7.2. AV* is large and positive for
(Me2N)2CO, and this is in accord with a fully dissociative mechanism
for this ligand exchange. AV* is negative for Me2SO but its value is
appreciably smaller, and this allows us to conclude that, while the mech-
anism is associative, there must be some tendency for the bound ligand
to move away from the cation and that the mechanism is, more pre-
cisely, association and interchange of ligands near the cation, which lim-
its the extent of the possible volume contraction were full association
to occur. Note also the difference between the meaning of volume
changes in this and the previous example. Here, the total volume before
and after reaction is unchanged, and we observe only the adjustment
necessary to allow reaction to proceed. There is thus no change in the
position of equilibrium with pressure, in contrast to what happens when
reaction produces a permanent volume change.
www.pdfgrip.com
Reaction monitoring of systems not at equilibrium 225
www.pdfgrip.com
226 NMR spectra of exchanging and reacting systems
— ^-—~ -rf
40 min 6h 22 h 46 h 54 h 78 h 102h
Figure 7.28 27
The 100 MHz ^-NMR spectra of the Al-Me protons with the
effect of the A1 removed by double irradiation of a 1:1 mixture of NaAlMe4
with NaAlEt4 in benzene with added (Me2N)3PO, as a function of time. (From
Ahmad et al (1984) Organometallics, 3, 389, copyright (1984) American
Chemical Society, reprinted with permission.)
www.pdfgrip.com
Reaction monitoring of systems not at equilibrium 227
(a)-
(b)
(c)
Figure 7.29 Three 40.5 MHz 31P spectra obtained with double irradiation of
the protons so as to simplify the spectra. The upper spectrum (a) is that of
the starting material and is distinguished by squares in all three spectra. The
central trace (b) was obtained 1.5 h later and shows the presence of another
substance in which the two phosphorus nuclei are still equivalent and which
is distinguished by triangles. After three days of reaction (c) both these spec-
tral patterns are much diminished and a third, more complex, pattern (x) has
taken their place. There is now a triplet of AB patterns, which shows that
the two phosphorus nuclei in the complex are no longer equivalent. (Example
supplied by Professor B.L. Shaw.)
www.pdfgrip.com
228 NMR spectra of exchanging and reacting systems
7+
www.pdfgrip.com
Reaction monitoring of systems not at equilibrium 229
[AI(OH)4]
Al) / ppm
Figure 7.31 The 104.2 MHz 27A1 NMR spectra of a partially hydrolysed aluminium salt solution as a func-
tion of time immediately following dilution with water. The 3+
aluminium concentration before dilution was
1.0 M and this was reduced to 0.1 M by dilution. [A1(OH2)6] is the chemical shift reference. A capillary
containing [A1(OH)4]~ was used to provide a concentration reference. (From Akitt and Elders (1988) /.
Chem. Soc., Dalton Trans., 1347, with permission.)
www.pdfgrip.com
230 NMR spectra of exchanging and reacting systems
7.3 QUESTIONS
www.pdfgrip.com
Questions 231
etc
www.pdfgrip.com
www.pdfgrip.com
Multiple resonance
and one-dimensional
pulse sequences
8
Until the advent of computer controlled NMR spectrometers, there
was a limited selection of NMR experiments available. However, with
computer control, it is possible to carry out a much wider selection of
experiments.
Many NMR experiments may be summarized by the sequence:
Relaxation - Preparation - Evolution - Mixing - Acquisition. During
the relaxation period, the nuclear spins are allowed to return towards
their Boltzmann equilibrium. In principle, the relaxation period should
be in excess of 57\ for the slowest relaxing signal. In practice, we are
impatient, and normally wait between Tl and 27\. It is during the
preparation period that the spins of the nuclei are tuned to give us
the information required. There is then an evolution period, which is
often related to a coupling constant. It is usually necessary to apply a
refocusing 180° pulse in the middle of the mixing period to make sure
that the resulting spectrum can be phased. One or more other pulses
are applied during the mixing period. At the end of the mixing period,
the FID is acquired, frequently with decoupling. In the more compli-
cated experiments, a string of pulses are applied, often to two or more
different nuclei. A single such experiment is a one-dimensional NMR
experiment. The data are collected in a single time dimension as the
FID and are then transformed into the frequency dimension to give
a conventional NMR spectrum. (Yes, strictly speaking there are two
dimensions, intensity and time or frequency, but such spectra are
described as one-dimensional, referring to the time dimension.) A
second time dimension is introduced by varying a time during the
mixing period. A series of FIDs are collected, each with a different
value of time, a technique which gives rise to two-dimensional NMR
spectroscopy. We will see in Chapter 9 that this provides a valuable
tool for assigning signals.
In the present chapter we will be examining a range of one-dimen-
sional NMR experiments. This is not intended to be an exhaustive
coverage, but many of the important techniques will be examined. In
all cases, the techniques give information on the connectivity between
nuclei in compounds, and hence on the structure of compounds.
www.pdfgrip.com
234 One-dimensional NMR spectroscopy
www.pdfgrip.com
Decoupling difference spectroscopy 235
(C)
(D)
(a) =
2.7 2.6 2.5 2.4 2.3 2.2
1
8( H)/ppm
Figure 8.1 A partial 400 MHz JH NMR spectrum of carvone in CDC13. (a)
The reference spectrum, where the decoupler was set at 8 10. (b) The spec-
trum where the decoupler was set on the signal at 8 6.77. (c) The difference
spectrum, where spectrum (a) is subtracted from spectrum (b). Responses are
observed only for the multiplets at 8 2.45 and 2.28 which have changed on
decoupling. The unchanged coupled spectrum gives the negative signals, while
the decoupled spectrum gives positive signals.
www.pdfgrip.com
236 One-dimensional NMR spectroscopy
(c>-
(b)-
(a)--'
7.8 7.6 7.4 7.2 7.0 6.8 6.6 6.4 6.2 6.0
1
H/ppm
www.pdfgrip.com
The nuclear Overhauser effect 237
(y-^PP
A transition X transition
6N ap—^ ^^pa8N
X transition A transition
(a)
^ aa 25N
A transition X transition
* X transition A transition
irradiate \, /
(b) ^ 'aa 1.56N
Figure 8.3 The energy level diagram for a homonuclear AX spin system, (a)
In the absence of irradiation, (b) In the presence of irradiation at the aa-a(3
X transition.
www.pdfgrip.com
238 One-dimensional NMR spectroscopy
w mfriV*2 5(5 + I K
W
°~ 120nV 1 + K - cos)2Tc2 ( }
w MtfriV*2 s(s + IK
Wl ( }
= SOnV 1 + o^V
where r is the distance between the two nuclei and a>T and oos are the
frequencies of the nuclei. The subscripts to W refer to the change in
the total spin quantum number when the transition described by W
occurs. The transition associated with W0 is a zero quantum transition,
that associated with W1 is a single quantum transition and that asso-
ciated with W2 is a double quantum transition.
The theoretical maximum NOE, iimax, when nuclei, /, are observed
and nuclei, /, are irradiated is given by
= . TfcC^-Wq)
www.pdfgrip.com
The nuclear Overhauser effect 239
When the extreme narrowing limit does not apply, r\max is dependent
on both frequency and TC (Fig. 8.5).
A qualitative understanding of the origin of the NOE can be under-
stood by examining the AX spin system with one of the protons
undergoing double irradiation (Fig. 8.6). The relative populations of
the energy levels follows from the Boltzmann distribution. Relative
to the pp energy level, the ap and Pa energy levels have an excess
population of 8N, while the aa energy level has an excess population
of 28N. When the decoupler is turned on at the X frequency, irradi-
ating both lines, transitions are induced between X spin states and the
populations of the connected energy levels are equalized. Hence
the populations of the pa and pp energy levels become M(0 + SN) =
M8N. Similarly, the populations of the ap and aa energy levels become
13
C{1H}
2 -,
1 -
o-
-1 --
3
l
-2-
-3-
-4-
1 5 1
N{H}
-5- i i i i i i i
-12 -11 -10 -9 -8 - 7 - 6 -5 -4
log(Tc)
Figure 8.5 The variation of Timax with correlation time for J, and
pH). The values have been calculated for B0 = 9.4 T.
www.pdfgrip.com
240 One-dimensional NMR spectroscopy
6N pa8N
(a)
irradiate
1.58N pa 0.58N
irradiate
(b)
Figure 8.6 The energy levels for a homonuclear AX spin system, (a) The
relative populations are given for each energy level at equilibrium, (b) The rela-
tive populations are given immediately after applying a radiofrequency
decoupling field at the X nucleus.
www.pdfgrip.com
The nuclear Overhauser effect 241
0
CO
D
Q.
decouple
Figure 8.7 The basic pulse sequence for the measurement of an NOE.
www.pdfgrip.com
242 One-dimensional NMR spectroscopy
www.pdfgrip.com
The nuclear Overhauser effect 243
(a)
www.pdfgrip.com
244 One-dimensional NMR spectroscopy
(e)-
(d)-
(c)-
(b)-
(a)-
8 7 6 5 4 3
1
H/ppm
Figure 8.9 The 400 MHz 1H NMR spectra of [(Ti5-C5H5)Mo(CHPhNMeCPh=NMe)(CO)2] in CDC13. (a)
The normal spectrum and (b) to (e), the NOE difference spectra with pre-irradiation at (b) 8 4.9, H12, rf-
C5H5, (c) 8 5.6, H8, (d) 8 5.5, H8/, and (e) 8 5.4, H12', Ti5-C5H5, with the shift scales progressively displaced
to the right to give a clearer picture. (Reproduced from Brunner et al (1983) /. Organomet. them., 243,
179, copyright (1983), with permission from Elsevier Science.)
www.pdfgrip.com
The nuclear Overhauser effect 245
Table 8.1 Maximum nuclear Overhauser effects for several pairs of nuclei
Irradiate >1i 19p
13 19 p 29 31 ! 13 19p
Observe *H C 15N Si P H C
0.5 1.99 -4.93 0.53 -2.52 1.24 0.47 1.87 0.5
i+^max 1.5 2.99 -3.93 1.53 -1.52 2.24 1.47 2.87 1.5
www.pdfgrip.com
246 One-dimensional NMR spectroscopy
(c)
(b)
(a)
240 220 200 180 160 140 120 100 80 60 40 20 0
13
C/ppm
Figure 8.10 The 100.62 MHz 13C spectra of carvone, dissolved in CDC13. (a) Obtained without proton irra-
diation, (b) With irradiation at 400.13 MHz with NOE suppression, (c) With irradiation at 400.13 MHz
and full NOE, otherwise the spectrometer conditions were identical. A large improvement in signal-to-
noise ratio is obtained with double irradiation, especially when the NOE is permitted.
www.pdfgrip.com
The nuclear Overhauser effect 247
decouple
'H
13, FID
d1 P1
Figure 8.11 The NOE suppression pulse sequence, dl is a relaxation delay of
at least 57\. pi is a 90° pulse. Decoupler timing to allow a fully decoupled
spectrum to be obtained without any distortion of intensity due to the NOE.
There will be a small build-up of NOE during the B2 pulse (typically 0.5 s)
and this must be allowed to die away completely before the next pulse. The
long delay time means also that the nuclear magnetization has decayed fully
before the next 90° pulse and there are no intensity distortions due to relax-
ation effects.
decouple
13, FID
d1 P1
Figure 8.12 The pulse sequence used to record a 1H coupled 13C NMR spec-
trum with full NOE. dl should be 5T{ or at least considerably longer than
the acquisition time. Decoupler timing is arranged to allow a coupled spec-
trum to be obtained but with the benefit of the full NOE increase in inten-
sities. There will be very little fall-off in NOE during the short time needed
to collect the FID data. This technique allows up to nine times reduction in
time over the basic method used to obtain a non-decoupled spectrum.
www.pdfgrip.com
248 One-dimensional NMR spectroscopy
(8.3)
(a)
200 150 100 50 0
(8.4) 13
C/ppm
13
Figure 8.13 The 100.6 MHz !
C NMR spectra of (8.4) in CDC13. (a) The 13C
NMR spectrum with no H irradiation, (b) The difference spectrum obtained
by pre-irradiating the proton at 8 2.37 at low power to develop an NOE in
the 13C NMR spectrum and subtracting spectrum (a), (c) An expansion of the
signals at 8 90.0 and 8 90.5 in (b). (Reproduced with permission from Aldersley
et al. (1983) /. Chem. Soc., Chem. Commun., 107.)
www.pdfgrip.com
The nuclear Overhauser effect 249
Figure 8.14 The dynamic NOE experiment. This is essentially a combination of the two experiments
described by Figs 8.11 and 8.12. The decoupler is first gated ON to allow the NOE to build up. The
amount of NOE increases if t is increased, reaching a maximum when t>9Tv At the end of time t, the
FID is produced by the 90° pulse and collected with continuing irradiation. This is removed when the
data collection is finished and the NOE allowed to decay to zero. Sufficient FIDs are collected to give
the required signal-to-noise ratio, and the experiment is repeated with different values of t, but always
the same number of FIDs are collected, so that spectral intensities can be compared directly.
www.pdfgrip.com
250 One-dimensional NMR spectroscopy
Figure 8.15 Fourier transform 13C spectra of biphenyl in CDC13 solution excited by 90° read pulses sepa-
rated by intervals of 300 s in order to ensure the re-attainment of equilibrium after each pulse. The spectra
are stacked as a function of the time t of proton irradiation prior to the application of the pulse. (From
Freeman et al. (1972) /. Magn. Reson., 7, 327, with permission.)
www.pdfgrip.com
The nuclear Overhauser effect 251
and so accounts for the small NOE. One can calculate that the part
of the relaxation which is purely dipolar, 3riDD, has the value 160s,
which implies a reduction of TC for the methyl group of around eight
times compared with the rest of the molecule.
A more complex and more informative example concerns the relax-
ation behaviour of the carbon nuclei in [Cr(CO)5(NC5H5)]. This is a
a-bonded complex in which one CO ligand in [Cr(CO)6] has been
replaced by an N-bonded pyridine molecule. The chemical interest in
such molecules arises because it is uncertain what sort of energy
barriers exist to the rotation of the Cr(CO)5 moiety relative to the
aromatic ligand. One way of probing the intramolecular motion is to
measure the relaxation rates of the 13C nuclei. Provided the mecha-
nism of relaxation is known, then correlation times of individual atoms
can be calculated. Two types of mechanism are expected to coexist in
such molecules: dipole-dipole for the protonated carbon atoms and
chemical shift anisotropy relaxation for the CO carbon atoms. The
first mechanism can be confirmed if the NOE is high, and low NOE
establishes its absence for the CO atoms. Another experiment is,
however, needed if we are unequivocally to establish the presence of
CSA relaxation, and this is achieved by measuring 7^ and T] at different
magnetic fields, though the rates of motion must be such that the
extreme narrowing condition is met at the higher field and spectro-
meter frequency
The Tv values were in this case obtained by a saturation-recovery
technique in which the 13C spin populations are equalized by the appli-
cation of a series of closely spaced 90° pulses and the recovery of the
signal intensity monitored as a function of time. The NOE was
determined by comparing signal intensities obtained with continuous
decoupling with those using gated decoupling of the type depicted in
Fig. 8.11. CDC13 was used as solvent and solutions were degassed
to ensure that there was no contribution to relaxation from solvent
nuclei or dissolved paramagnetic oxygen. The two spectrometer
frequencies used were 125.7MHz (11.7 T, 1H resonates at 500 MHz)
www.pdfgrip.com
252 One-dimensional NMR spectroscopy
Table 8.3 Relaxation and NOE data for the 13C nuclei in two chromium carbonyl complexes
www.pdfgrip.com
One-dimensional multipulse sequences 253
the two halves of the complex are rotating independently around the
N-Cr bond, with the rates of motion determined primarily by inter-
actions of the groups with the solvent. The data for the (T]6-arene)
complex, [(7i6-C6H6)Cr(CO)3], are also given in Table 8.3. They
resemble closely the previous set of data, though the correlation times
calculated are much shorter and the speed of rotation of the benzene
is high indicating its very small interaction with the solvent.
13C
d1 p1 d2 p2 d2
Figure 8.16 The pulse sequence for the / modulation pulse sequence, dl is
for relaxation, pi is a 90° pulse, d2 is 1/J(13C,1H), and p2 is a 180° refocusing
pulse.
www.pdfgrip.com
254 One-dimensional NMR spectroscopy
a 13C signal with no 1H coupling (Fig. 8.18(a)). As the 13C NMR signal
is on resonance, it does not matter how long we wait after the initial
90° pulse as the signal will remain with its original alignment. However,
in a normal 13C NMR spectrum, many of the signals will be away from
resonance, and will precess relative to the rotating frame at different
frequencies. Thus, by the time we collect the FID, they will be very
much out of phase and enormous phase corrections will be needed to
rephase the spectrum. This is avoided by placing a 180° refocusing
pulse in the middle of the sequence so that nuclei are all in phase
again when the collection of the FID is initiated, see section 6.3. This
is a general feature of these pulse sequences.
Let us now consider a 13CH group, again on resonance (Fig. 8.18(b)).
The rotating frame is at the frequency of the centre of the 13CH
doublet. One line of the doublet will be +J/2 Hz away and the other
will be -JI2 Hz away. The result is that the 13C nuclei attached to 1H
nuclei with a or p spins will be precessing at rates different from that
of the rotating frame. After 1/27 s, one group will have precessed +90°
www.pdfgrip.com
One-dimensional multipulse sequences 255
Figure 8.18 The behaviour of the nuclear spins in the x'y' rotating frame
during the /-modulation pulse sequence. In order to simplify the picture, it
is assumed that the 13C NMR signal is on resonance, (a) A 13C not bearing
1
H. (b) A 13CH group, (c) A 13CH2 group, (d) A 13CH3 group. The 180° pulse
and second 1/J wait are omitted as they follow from the refocusing pulse
description in section 6.3.
and the other -90°. After l//s, they will have processed ±180° and
refocused. The decoupler is now switched on and the doublet collapses
to a singlet, which precesses at the same frequency as the rotating
frame. The signal from the 13CH group is 180° out of phase compared
with the non-proton bearing 13C.
www.pdfgrip.com
256 One-dimensional NMR spectroscopy
www.pdfgrip.com
One-dimensional multipulse sequences 257
13C FID
d1 p1 d2 p2 d2 d3 p2d3
Figure 8.19 The APT pulse sequence, dl is a relaxation delay, pi is a 45° pulse, d2 is I//, p2 are 180°
pulses, and d3 is the pre-acquisition delay.
8.3.2 INEPT
There is a family of pulse sequences which gain an enhancement in
signal intensity from a higher frequency coupled nucleus, usually 1H.
Equation (1.3) gives the population difference between the upper and
lower energy levels for an / = 1/2 nucleus. The population excess in
the lower energy level is proportional to jjuj or ylt Hence for :H, with
yH = 26.7510 x 107 rad T'1 s-1, the population excess is approximately
four times greater than for 13C, with yc = 6.7263 x 107 rad T'1 s'1. If it
were possible to transfer this population excess from !H to 13C, then
the 13C NMR signal strength would increase by a factor of approxi-
mately four times. This can be done, but the situation is not all win
as the NOE is lost. Despite this, population transfer is generally
preferred as the gains, y$lyi, are normally greater than those from NOE,
1 + 7s/27j, see Tables 8.1 and 8.2, while the uncertainty associated with
NOE, that dipole-dipole relaxation is required to be dominant, is
removed. The gain is spectacular for very low frequency nuclei such as
103
Rh, especially as in this case NOEs are not normally observed.
The principle behind population transfer can be illustrated for
an AX system consisting of a 13C1VL pair of spins (Fig. 8.20). The
relative populations of the energy levels are given. Before we apply a
www.pdfgrip.com
258 One-dimensional NMR spectroscopy
selective 180° pulse, the population differences for the two 1H transi-
tions are both 48N, while for the two 13C transitions they are both 8N,
in Fig. 8.20(a). In Fig. 8.20(b), after applying the selective 180° pulse
to one !H transition, the population differences, and hence intensities
of the 13C transitions have changed to 58N: (-38N). Hence one tran-
sition has increased in intensity by 48N units and the other decreased
by -48N units. If we were to now repeat the measurement, applying
the selective 180° !H pulse to the other !H transition, we would get a
13
C doublet with intensities (-38N): 58N. If these signals are now
subtracted, the result is a 88N: (-88N) doublet compared with a
28N : 28N doublet if a simple 13C NMR spectrum had been taken using
two acquisitions. The gain has been a factor of four, or more accu-
rately VYc-
This experiment is impractical to carry out routinely as it would be
necessary to carry it out for each separate carbon environment in
the molecule. However, exactly the same result can be achieved
by the pulse sequence in Fig. 8.21 for all the 1H coupled to X in a
molecule, where X is an / = 1/2 nucleus. This is the Insensitive Nuclei
Enhanced by Polarization Transfer, or INEPT, pulse sequence.
(a)
(b)
Figure 8.20 The effect of applying a 180° selective pulse to one line of the
!H AX doublet of a 13CH group, (a) Before, (b) After.
www.pdfgrip.com
One-dimensional multipulse sequences 259
'H
d1 p1 d2 p2 d2 p3
X.
P4 P5
Let us examine the effect of these pulses on the !H nuclei (Fig. 8.22).
We need to keep track of both the !H and X nuclei. In order to do
this, 1VL nuclei attached to X nuclei with a-spin are labelled a, and
those attached to X nuclei with p-spin are labelled b. The first (90°)r
pulse brings the XH magnetization into the xy plane pointing in the y'
direction. If the !H nuclei are on resonance, the two lines of the
doublet rotate at ±/XH/2 Hz- After 1/4^xn s > theY wili have rotated ±45°.
A 1H(1SQ°)X refocusing pulse is applied. If nothing else were to be
done, the two lines of the doublet would refocus in the -y' direction
after a further l/4/XH s. Instead a X(180°) pulse is also applied. This
has the effect of exchanging the X spin states. The protons which were
attached to X with a-spin are now attached to X with (3-spin and vice
versa. The result is that the !H nuclei precession direction in the
rotating frame is reversed, and after the further l/4/XH s they end up
aligned 180° with respect to each other. The final lH(90°)y pulse rotates
Figure 8.22 The effect of the INEPT pulse sequence on the !H magnetiza-
tion in the xy plane. The diagrams start following the 1H(9Q°)X pulse.
www.pdfgrip.com
260 One-dimensional NMR spectroscopy
their magnetization back into the z-direction. One line of the doublet
is pointing in the +z direction, while the other is pointing in the -z
direction. This has achieved the same effect as the selective 180° pulse
described above, but has produced the effect for all the !HX spin
systems in the sample. The X(90°) pulse then produces an X FID with
polarization transfer. Phase cycling is used to alternately invert the
lines of the !HX doublet. For example, if instead of using a lH(9Q°)y
pulse as the final 1H pulse, a lH(9Q°)_y pulse was used, the phase of
the 1H doublet is reversed. Phase cycling of the receiver permits
addition or subtraction of the spectra. The resulting 1H coupled X
spectrum does not have the usual 'Pascal triangle' intensities, but
rather those given in Table 8.5. This is illustrated in Fig. 8.23 for
some rhodium hydrides. Note that to distinguish between a !:(-!)
doublet and a 1:0: (-1) triplet, the separation of the lines has to be
measured, and is / in the former case and 2J in the latter, with / being
pre-determined from the 1H NMR spectrum.
In order to be able to *H decouple, the pulse sequence has to be
extended. This is because in a !H coupled INEPT spectrum there
is as much positive intensity as negative intensity. The result of
decoupling is to give zero intensity. Fortunately the answer is very
straightforward. A delay is placed between the basic INEPT pulse
sequence of Fig. 8.21 and acquisition (Fig. 8.24). Of course, it is neces-
www.pdfgrip.com
One-dimensional multipulse sequences 261
Table 8.5 The INEPT Pascal triangle. It is constructed in the same way as
the normal Pascal triangle adding together the pair of numbers on the line
above, except that it is extended by putting 1 and -1 at the ends of the
next line.
0
-1 1
0 1 -1
1 1 - 1 - 1
1 2 0 - 2 - 1
1 3 2 - 2 - 3 - 1
1 4 5 0 - 5 - 4 - 1
1 5 9 5 - 5 - 9 - 5 - 1
1 6 14 14 0 -14 -14 -6 -1
1 7 20 28 14 -14 -28 -20 -7 -1
etc.
sary to include a refocusing 180° pulse for both !H and X, p4 and p7,
halfway through the delay.
Figure 8.25 shows the response of the X-magnetization in the xy
plane for an AX group. The first diagram shows the state of the magne-
tization following the X(90°);c pulse, p6. The two lines of the X doublet
are 180° out-of-phase. It is necessary again to follow the spin states
of the other nuclei. The X-nuclei attached to XH nuclei with a-spin
are labelled a and those attached to !H nuclei with (3-spin are labelled
b. After 1/4/ s, the magnetizations have rotated ±45° in the rotating
frame, assuming again that the X signal is on resonance. As usual it
is necessary to use an 180° X refocusing pulse. It is then necessary to
use a 180° !H pulse to reverse the 1H spin states and hence the labels
on the magnetization vectors. The result is that they continue to
precess in the same direction and refocus after a further 1/4/s. The
1
H decoupler is then turned on and the 1H decoupled FID collected,
enhanced by polarization transfer.
www.pdfgrip.com
262 One-dimensional NMR spectroscopy
Figure 8.25 The effect of the INEPT pulse sequence on the X magnetization
in the xy plane.
8.3.3 DEPT
DEPT is the pulse sequence of choice to edit 13C NMR spectra. The
pulse sequence is given in Fig. 8.28. It is not possible to use a simple
vector model to describe this pulse sequence. The 13C editing is carried
out by choosing a suitable p3. Values of 45°, 90°, and 135° are used.
With a 45° pulse, the CH, CH2, and CH3 signals are all positive, a 90°
pulse only gives CH signals, and the 135° pulse gives CH and CH3
positive, while the CH2 signal is negative. Suitable addition and
www.pdfgrip.com
One-dimensional multipulse sequences 263
1.2
0.8
0.4
&
g 0
o>
J:
-0.4
-0.8
-1.2
0.125 0.25 0.375 0.5
d3 in units of -j- s
j
Figure 8.26 The response of CH, CH2, and CH3 groups as a function of d3
when the refocused INEPT pulse sequence is used.
22 21 20 19
13
C/ppm
Figure 8.27 A partial 100.62 MHz 13C NMR spectrum of one of the methyl
groups of carvone in CDC13. The spectrum was recorded using the INEPT
pulse sequence and shows !H coupling. The -1 : -1:1 :1 pattern arises
from the INEPT pulse sequence which was set up for a J(13C1H) - 145 Hz
corresponding to an average ^(^CH). The doublet of doublet of doublets
multiplet structure of each of the four lines arises from the much smaller
3 13
/( C?H).
www.pdfgrip.com
264 One-dimensional NMR spectroscopy
'H
d1 p1 d2 p2 d2 p3
13C
P4 p5 d2
pulse of variable length, see text, p4 is a 90° X pulse and p5 is a 180° X pulse.
i£H3
CH2
CH
150 100 50 0
5(13C)
Figure 8.29 Edited 13C DEPT spectra of ristocetin at 100.6 MHz. The lower spectrum is the normal broad-
band decoupled spectrum showing all the carbon resonances. Note that even the resonance of the CD2H
groups of the solvent dimethylsulfoxide-d6 appears in its appropriate trace. (From Bruker, with permission.)
8.3.4 PENDANT
Although DEPT is an excellent pulse sequence to use for the
observation of CH, CH2, and CH3 groups, it does not detect non-
protonated 13C atoms. / modulation or APT used to be the initial
www.pdfgrip.com
One-dimensional multipulse sequences 265
'H
p1 p2 p3 p4
X
dl p5 d2 p6 d2 P7 d3 p8 d3 '
www.pdfgrip.com
266 One-dimensional NMR spectroscopy
C/CH2
CDCI3
CH/CH3
- i | i | i | i | i | i | i | i | i | i | i | i | i |
decouple
13C
d1 p1 d2 p2 d2 P3 d3 p4
www.pdfgrip.com
Exercises in spectral interpretation 267
'H decouple
13C
d1 p1 d2 p2 d2 P3 d3 p4 d2 p2 d2
(b)-
(a)
n i r
5910 5860 3600 3550 3040 2990
Hz
Figure 8.34 The application of the one-dimensional INADEQUATE pulse sequence to determine
/(13C13C) in [(T]5-C5H5)Ni(l,3,4-Ti3-2,2-dimethylbutenyl)] at 100.62MHz in d8-THF. (a) The signals due to
C3, C2, and C6. (b) INADEQUATE NMR spectra with d2 - 0.0062 s corresponding to /(13C13C) - 40 Hz.
(c) INADEQUATE NMR spectra with d2 - 0.08 s corresponding to /(13C13C) - 3 Hz. (Reproduced from
Benn and Rufinska (1982) /. Organomet. Chem., 238, C27, copyright (1982), with permission from Elsevier
Science.)
www.pdfgrip.com
268 One-dimensional NMR spectroscopy
We have shown in the previous chapters that the carbon spectra are
capable of yielding much information through a series of quite time-
consuming experiments. These facilities are normally reserved for the
more difficult samples, and often it will be sufficient to have simply
the broad-band proton-decoupled spectrum. This gives the chemical
shift information, details of coupling to nuclei other than hydrogen,
and a carbon count, provided all the likely errors in line intensity are
taken into account. It is also certain that the proton spectrum will be
available, since this is so easy to obtain, and interpretation will
be based on the two sets of data taken together. The carbon spectra,
of course, give information about atoms not bonded to hydrogen,
which is not available in the proton spectra.
The carbon chemical shifts are much more widely dispersed than
are the proton shifts. The ranges within which different types of carbon
atom resonate are shown on the chart in Fig. 8.35. The chemical shift
of a given carbon atom in a family of compounds is sensitive to the
influence of all four substituents, and for alkanes, for instance, can be
predicted using the Grant and Paul rules
8; = -2.6 + 9.1na + 9.4/10 - 2.5ny + 0.3na
Tl\US
ni-uv/i ^^H 1
vi ry*, ^^^^M
cyclopropanesl
CH3R 1•
C BpR2™
Ch«R3"¥
CHsHal 1 •
r.Hj\i^«
'*•' • vy " "
•
CHp •
QP.M^n ^^H
1 IVXI l^\.X
••
R2CHO •
p_/^n
~ ^
ig^v^ -f-
RCH2S 1 •I
Alkynes •l
Alkenes ^" •1
1'CH ••
Aromatics < CN
•
.CO" •
i—»/-\x-\ 1ijmm•
i-iu U2' ^^H
RCHOI
R^CO •1
•
M(CO)
••
www.pdfgrip.com
Exercises in spectral interpretation 269
1.3 1.2
9 8 7 6 5 4 3 2 1 0
1
(a) H/ppm
Exercise 4 (a) The 400.13 MHz 'H NMR spectrum of a solution of ethyl benzene in C6D6.
www.pdfgrip.com
270 One-dimensional NMR spectroscopy
N=N N N-
Exercise 5 The 400 MHz 1H NMR spectrum (upper) and 100.62 MHz 13C{1H}
JMOD NMR spectrum (lower) of the aromatic compound C12H8N2. The 13C
NMR spectrum shows signals at 8 143.3,130.4, and 129.6 due to the compound
and 77.0 due to CDC13. The proton spectrum is diagnostic of an [AB]2 four-
spin system in which there are two pairs of protons with the same chemical
shift 8A, 5B, but different coupling constants /(AB), /(AA'), /(AB') and
7(BB'). Only three lines are observed in the carbon spectrum. This informa-
tion is sufficient to identify the compound, some suggestions for which appear
below the spectra. Example supplied by A. Romer.
www.pdfgrip.com
Exercises in spectral interpretation 271
www.pdfgrip.com
272 One-dimensional NMR spectroscopy
www.pdfgrip.com
Two-dimensional NMR
spectroscopy 3G
Two-dimensional NMR spectroscopy has become an important tool
for chemists. Several experiments are possible and new variants are
being continually invented, such is the power of the method. The two
dimensions are dimensions of time as previously discussed at the begin-
ning of Chapter 8. One of these is already familiar, and is the time
domain within which we collect the FID output from the spectrom-
eter and which contains frequency and intensity information. The
second dimension refers to the time elapsing between the application
of some perturbation to the system and the onset of the collection of
data in the first time domain. This second time period is varied in a
regular way and a series of FID responses are collected corresponding
to each period chosen.
In order to illustrate how two-dimensional NMR spectroscopy
works, we will show how the one-dimensional / modulation described
in section 8.3.1 can be converted into a two-dimensional experiment.
The delay d2, which is related to /, is replaced by a variable delay
fj/2, where ^ is made to assume many values. The pulse sequence is
shown in Fig. 9.1 which is essentially that of Fig. 8.16 though we should
note that in practice it is more common for the decoupler to be off
during the first delay and on during the second, though this makes no
difference to the outcome. Examination of the behaviour of the spin
magnetization in Fig. 8.18(b) shows that the intensity of the 13C signal
from a CH entity is a function of the delay time and varies as
cos(2ffrtlIJ). It will also decay by the T2 process though the 180° pulse
refocuses the magnetization and this is the true T2\ see section 6.3.
Similarly, the 13C signal of a CH2 group will be seen from Fig. 8.18(c)
to remain always positive and to follow a [cos(2(Trr1//) + 1] law. In the
two-dimensional form of the experiment, a series of FIDs are collected
corresponding to different values of ^. Each FID is then Fourier trans-
formed individually to give a series of spectra, one for each value of
*! chosen. So we have spectra varying as a function of/ 2 (corresponding
to the t2 dimension) and covering a range of ^ values as in Fig. 9.2.
If we look at the intensity at 8 42.45, which arises from a CH, we
will observe that it changes across the spectra as a decaying cosine
wave. This is made clearer by the section through the spectra
shown in Fig. 9.3. The frequency of this wave gives the C-H coupling
constant. The signal is not however, a simple decaying cosine wave
www.pdfgrip.com
274 Two-dimensional NMR spectroscopy
decouple decouple
13C FID
d1 p1 A. p2 A.
2 2
Figure 9.1 The J modulation pulse sequence, dl is for relaxation, pi is a 90°
pulse, f/2 and t2 are the two time dimensions for the two-dimensional exper-
iment, and p2 is a 180° pulse.
0.040 s
0.035 s
0.030 s
0.025 s
0.020 s
0.015s
0.010s
0.005 s
0.001 s
543.10 6 42.45
Figure 9.2 The individual spectra of a partial 7-resolved 100.62 MHz 13C NMR spectrum of carvone in
CDC13 obtained using the pulse sequence in Fig. 9.1. The individual FIDs have been transformed in the
r2//2 direction and plotted for ^ values from 0.001 to 0.04 s, incremented by 0.001 s, given on the right.
Note the decaying cosine oscillations of each signal in the fj//i direction.
www.pdfgrip.com
Two-dimensional NMR spectroscopy 275
• ' • i • • • i • < • i • > • i • - • i ' < • i • i ' i • ' ' i • ' • i ' ' ' i • i • i —
200 180 160 140 120 100 80 60 40 20 0
13
C/ppm
Figure 9.4 The individual spectra of a /-resolved 100.62 MHz 13C NMR spectrum of carvone obtained
using the pulse sequence in Fig. 9.1. The individual FIDs have been transformed in the t2/f2 and ^//i direc-
tions and plotted for a selection of ^ values.
www.pdfgrip.com
276 Two-dimensional NMR spectroscopy
-100
- -50
OHz
- 50
- 100
Figure 9.5 The contour plot of a /-resolved 100.62 MHz 13C NMR spectrum
of carvone in CDC13 obtained using the pulse sequence in Fig. 9.1.
www.pdfgrip.com
Two-dimensional NMR spectroscopy 277
Figure 9.6 Cross-sections of the signals at (a) 8 42.45 and (b) 43.10 of a /-resolved 100.62 MHz 13C NMR
spectrum of carvone in CDC13 obtained using the pulse sequence in Fig. 9.1, showing the multiplicity due
to !H coupling. Note that due to the pulse sequence used the coupling constants have been halved.
OH
(a) (b) 10
www.pdfgrip.com
278 Two-dimensional NMR spectroscopy
FID
d1 p1 tl P2 ll
2 2
Figure 9.8 The /-resolved pulse sequence, dl is for relaxation, pi is a 90°
pulse, tJ2 and t2 are the two time dimensions for the two-dimensional exper-
iment, and p2 is a 180° pulse.
www.pdfgrip.com
Two-dimensional NMR spectroscopy 279
-15
-10
-5
10
15
Hz
2.6 2.5 2.4 2.3
1
H/ppm
Figure 9.9 A contour plot of the /-resolved 400 MHz 1H NMR partial spec-
trum of carvone in CDC13 before tilting. A projection of the spectrum is shown
above.
www.pdfgrip.com
280 Two-dimensional NMR spectroscopy
-15
-10
-5
10
15
Hz
2.6 2.5 2.4 2.3
1
H/ppm
Figure 9.10 A contour plot of the /-resolved 400 MHz !H NMR partial spec-
trum of carvone in CDC13 after tilting. A projection of the spectrum is shown
at the top.
trace. There are, for instance, two 1:1:1:1 quartets at 8 4.6 and 3.73,
and we can see immediately that the former consists of two equal over-
lapping doublets whereas the latter is a doublet of doublets involving
only one chemical shift value. The multiplet structure appears in the
coupling dimension. 128 values of T were used and 128 FIDs were
obtained, which limits the digital resolution to 0.39 Hz. It is a simple
matter to obtain all the coupling constants to within this limit. What
can we deduce about the spectrum of D-amygdalin from this chart?
The lines are already assigned on the figure but we will attempt to
proceed as if this had not already been done. Four lines are marked
as originating from proton-6. They are all doublets of doublets, but
one of the couplings is rather large at about 11 Hz and is found
nowhere else in the spectrum, and we can think immediately in terms
of a possible geminal pair. The two large couplings are slightly
different, sufficiently so that it is possible to separate the four reso-
nances into two pairs linked by identical coupling constants. These are
then the CH2 protons linked to the two rings, although we do not actu-
ally know yet which is which. Each proton of these geminal pairs is
further coupled to proton 5, though the responses from both of these
www.pdfgrip.com
Two-dimensional NMR spectroscopy 281
•15
•10
-5
10
15
Hz
2.6 2.5 2.4 2.3
1
H/ppm
Figure 9.11 A contour plot of the /-resolved 400 MHz 1H NMR partial spec-
trum of carvone in CDC13 after tilting and symmetrizing about a horizontal
running through the middle of the spectrum at 0 Hz. A projection of the spec-
trum is shown at the top.
www.pdfgrip.com
282 Two-dimensional NMR spectroscopy
Figure 9.13 The normal, i.e. one-dimensional, 500 MHz !H NMR spectrum
of D-amygdalin. There is a pattern in 2 : 3 intensity ratio at 7.6 ppm due to
phenyl, a CHCN singlet at 5.88 ppm, HOD at about 4.75 ppm and a complex
pattern to low frequency due to the remaining 14 sugar protons. (From Ribiero
(1990) Magn. Reson. Chem., 28, 765-7, copyright (1990) John Wiley and Sons
Ltd, reprinted with permission.)
COSY stands for Correlation SpectroscopY, and was the first two-
dimensional technique to be proposed by Jeener in 1971. As we shall
see, it has given rise to many variants, which depend upon the exis-
tence of spin-spin coupling between nuclei to provide supplementary
responses that relate the chemical shift positions of the coupled nuclei.
It is equivalent to carrying out simultaneously a series of double-reso-
nance experiments at each multiplet in the spectrum and looking for
the part of the spectrum where a perturbation has occurred. It is thus
rapid and avoids the need to know what irradiation frequency to use
www.pdfgrip.com
Homonuclear COSY NMR spectroscopy 283
10
10
Hz
4.8 4.6 4.4 4.2 4.0 3.8 3.6 3.4 3.2
I 11 I /~ ~.
www.pdfgrip.com
284 Two-dimensional NMR spectroscopy
FID
d1 p1 * P2
Figure 9.15 The basic pulse sequence for the COSY experiment, dl is a relax-
ation delay, pi is a variable pulse, fixed for a given experiment, p2 is a 90°
pulse, and ^ is the variable delay for the second dimension.
axes and the spectrum along the diagonal, which projects onto either
axis as the normal spectrum. The distortion introduced by the pulse
sequence introduces extra responses, off the diagonal, for those
protons which are spin coupled and none for those which are not.
These extra resonances appear on the two-dimensional map at the
points which describe the two chemical shifts of the coupled protons
and so enable the shifts of coupled resonances to be discovered, or
correlated. These peaks are often called cross-peaks.
A partial COSY-90 spectrum of D-amygdalin is shown in Fig. 9.16,
concentrating on the low frequency part of the spectrum where all the
interest lies. The diagonal runs from top left to bottom right of
the map and when projected onto a chemical shift axis is seen to
be the one-dimensional spectrum, in this case shown at the top of the
figure. Several off-diagonal responses are evident, disposed symmetri-
cally about the diagonal, and it is these we must examine for infor-
mation about correlation. The anomeric signals at about 4.6 ppm can
be used as a point of entry for analysis of the spectrum, and the connec-
tivities to the H2 signals can easily be traced. Unfortunately, the diag-
onal is very cluttered in the 3.4 to 3.6 ppm region, and we cannot move
on with any degree of certainty to the connectivities to protons further
around the ring. This is a disadvantage of COSY-90, which we will
see how to circumvent shortly. The doublet of doublets at 4.22 ppm
has a cross-peak to the complex multiplet at 3.92 ppm and to the multi-
plet at 3.62 ppm, the latter also being correlated with the 3.92 ppm
resonance. This pattern is that of an ABX spin system and so has to
be due to H6, H6' and H5 of one ring. The multiplet at 3.62 ppm is an
octet and so has to be H5, in agreement with the /-resolved data, which
has already allowed us to locate the protons 6. It follows from the
cross-peak near the diagonal that H4 of this ring is at about 3.5 ppm.
A second, similar pattern, connects signals at 3.9, 3.75 and 3.49 ppm
and presumably arises from the protons 5 and 6 of the second ring.
H4 cannot be assigned in this case. It remains to assign this and the
H3 protons.
Figure 9.16 shows the spectrum before symmetrization. The
symmetrized spectrum is in Fig. 9.17.
www.pdfgrip.com
Homonuclear COSY NMR spectroscopy 285
HOD
5.0 4.8 4.6 4.4 4.2 4.0 3.8 3.6 3.4 3.2
1
H/ppm
www.pdfgrip.com
286 Two-dimensional NMR spectroscopy
5.0 4.8 4.6 4.4 4.2 4.0 3.8 3.6 3.4 3.2
1
H/ppm
2
B5B9 B B11B3B4B6B10 B1
-35
-30
-25
-20 Q.
Q.
-15,23
-10
-5
(9.1) 0
0 -5-10-15 -20-25 -30-35
11
B/ppm
Figure 9.18 The 128.4 MHz nB COSY-90 NMR spectrum of 9-SMe2-7,8-mdo-
C2B9Hn, (9.1), in CDC13. Numbers are used to identify the boron atoms and
• to indicate the carbon atoms. (Reproduced from Rosair et al (1997) /.
Organomet. Chem., 536, 299, copyright (1997), with permission from Elsevier
Science.)
www.pdfgrip.com
Homonuclear COSY NMR spectroscopy 287
5.0 4.8 4.6 4.4 4.2 4.0 3.8 3.6 3.4 3.2
1
H/ppm
www.pdfgrip.com
288 Two-dimensional NMR spectroscopy
spectrum shows that the off-diagonal signals are not present. The result
is that the spectrum close to the diagonal is also much cleaner, and it
is easier to identify cross peaks in crowded regions of the spectrum,
for example between 8 3.3 and 3.6, compare Figs 9.16 and 9.19.
Examination of the cross peak at 8 3.5/3.9 shows that in the COSY-
90 NMR spectrum the responses map out a rectangle (Fig. 9.20(a)),
while in the COSY-45 NMR spectrum, half the responses are missing
(Fig. 9.20(b)). The result is that the cross peak appears to 'lean'. This
provides us with useful information. If we compare the two B6 cross
peaks at 8 3.5 and 3.74, they lean in opposite directions. This is because
the cross peak at 8 3.74 is to protons B6' and B5.2/(B6,B6') is a geminal
coupling constant and is negative, while 3/(B6,B5) is a vicinal coupling
3.2
3.4
3.6 _
5
•o
3
3.8
4.0
4.2
4.4
1 1
(a) H/ppm (b) H/ppm
Figure 9.20 Partial 400 MHz 'H (a) COSY-45 and (b) COSY-90 NMR spectrum of D-amygdalin in D2O.
The active coupling constant is indicated by dotted lines and arrows. The diagonal resonances are near
8 3.95.
www.pdfgrip.com
Homonuclear COSY NMR spectroscopy 289
constant and is positive. The direction of 'lean' gives the relative sign
of the coupling constant that is operating. The missing peaks also tell
us which coupling constant is operating between a pair of protons. It
is the one where both components are present.
5.0 4.8 4.6 4.4 4.2 4.0 3.8 3.6 3.4 3.2
1
H/ppm
www.pdfgrip.com
290 Two-dimensional NMR spectroscopy
4.0 3.5
1
H/ppm
Figure 9.22 A cross-section through a signal at 8 3.74 fron the partial phase-
sensitive 400 MHz *H COSY spectrum of D-amygdalin shown in Fig. 9.21.
The cross-section is through B 6 , which gives the out-of-phase signal at 8 3.74,
and shows the coupling to B6 at 8 3.94 and B5 at 8 3.49.
www.pdfgrip.com
Heteronuclear COSY NMR spectroscopy 291
cpd
d1 p1 jj_ f| d2 p2 d3
2 "2
FID
(a) P3 P4
'H cpd
d1 p1 p5 d4 p6 d5 p7 d2 p2 d3
13C FID
P3 P4
Figure 9.23 Pulse sequences used to produce ^-^C hetero COSY spectra, pi, p2, and p5 are (90°),
pulses, p7 is a (90)% pulse, p4 is a 90° pulse, p6 is a (180°), pulse and p3 is a 180° pulse, dl is the relax-
ation delay, ^ is the time for the second dimension, (a) Upper sequence is the one used for detecting
correlations through 1J(13C1H) as used for Fig. 9.24. d2 and d3 are 1/2J for ^("CH) and2 0.0036 s was
used, (b) The lower sequence is used to produce Fig. 9.25 where the correlation is through J(13C1H) and
3 13 1
J( C H). d2 and d3 are 1/2J for 2*J(13C1H) and 0.14 s was used. d4 and d5 are 1/2J for 1J(13C1H)9 0.0036 s.
cpd represents composite pulse decoupling.
www.pdfgrip.com
292 Two-dimensional NMR spectroscopy
-3.5
-4.0
-4.5
HOD
-5.0 3f
"D
-5.5 I
CH- -6.0
-6.5
-7.0
m,p -7.5
130 120 100 90 80 70 60 50
13
C/ppm
Figure 9.24 100.62MHz 13C-1H NMR correlation plot for hydrogen and
carbon atoms that are directly bonded in the molecule D-amygdalin in D2O.
www.pdfgrip.com
Heteronuclear COSY NMR spectroscopy 293
www.pdfgrip.com
294 Two-dimensional NMR spectroscopy
d1 p1 ?1 p2 p3 p2
Figure 9.26 The basic pulse sequence for the HOHAHA or TOCSY experi-
ment, dl is a relaxation delay, pi is a 90° pulse, ^ is the variable delay for
the second dimension, p2 is a pulse, typically 2.5 ms, attenuated to the spin
lock level, and p3 is the spin lock pulse sequence, where the transmitter power
has been attenuated so that the 90° pulse is of the order of 40 JJLS.
www.pdfgrip.com
HOHAHA or TOCSY 295
3.5
4.0 f
3
4.5
Figure 9.28 Two cross-sections from the 400 MHz 1H HOHAHA or TOCSY experiment shown in Fig.
9.27. (a) A cross-section though 8 4.58. (b) A cross-section through 8 4.63. Note the two separate sets of
signals for the protons of the two sugar residues.
www.pdfgrip.com
296 Two-dimensional NMR spectroscopy
d1 p1 d2 p2 d2 p3 p4
Figure 9.29 The INADEQUATE pulse sequence, dl is for relaxation, pi, p3,
and p4 are 90° pulses, d2 is 1/47', where J is the coupling constant between
the nuclei it is wished to detect, usually, 1J(13C13C), tl9 is the second dimen-
sion for the two-dimensional experiment, and p2 is a 180° pulse.
www.pdfgrip.com
Overhauser and magnetization transfer based spectroscopy 297
(9.2)
4.61 ppm, and this is then in the A ring, as, conveniently, we have
labelled the peaks throughout. The inverse experiment is also possible,
in which the doublets at 4.61 and 4.58 ppm are irradiated in turn and
the effect observed on the CHCN proton. This leads to the same
conclusion, though with less clarity, since it is not possible to irradiate
such close resonances with sufficient selectivity so that one gives an
NOE and the other none. A weaker effect is detected in the latter
case due to spill-over of the irradiation power into the 4.61 ppm
doublet region.
It is, of course, possible to carry out a two-dimensional version of
this experiment on D-amygdalin. Such experiments are very important
in the case of large, flexible biomolecules such as peptides. In solu-
tion, it is possible, once the proton resonances of identifiable residues
have been assigned, to determine which are in close proximity in space.
Thus the way the chains of such molecules are folded can be ascer-
tained, and the data currently being obtained in solution studies of
large molecules are comparable in accuracy with crystallographic data
on the same molecules. This experiment is called Nuclear Overhauser
www.pdfgrip.com
298 Two-dimensional NMR spectroscopy
--1000
- -500
f. *
- 500
1000
500 -500 Hz
9.6.1 NOESY
The NOESY (Nuclear Overhauser Enhancement SpectroscopY)
experiment is the two-dimensional version of NOE difference NMR
spectroscopy, see section 8.2. The NOESY pulse sequence is given in
Fig. 9.32.
The NOE builds up during the mixing time d2. This presents
an experimental problem. The individual contributions to 7\ from an
www.pdfgrip.com
Overhauser and magnetization transfer based spectroscopy 299
FID
d1 P1 t\ p2 d2 p3
Figure 9.32 The NOESY pulse sequence, dl is the relaxation delay, pi, p2,
and p3 are 90° pulses, ^ is the time for second dimension, and d2 is the mixing
delay for the NOE to build up.
7 6 5 4 3
1
H/ppm
www.pdfgrip.com
300 Two-dimensional NMR spectroscopy
Figure 9.34 128.37 MHz nB EXSY NMR spectrum of a 1.05 :1.0 M mixture
of BC13 and BBr3 at 400K using a mixing time of 0.05 s. (Reproduced with
permission from Derose et al (1991) /. Magn. Reson., 93, 347.)
www.pdfgrip.com
Overhauser and magnetization transfer based spectroscopy 301
178-
179-
180-
181 -
182-
c 183-
£184-
o: 185-
£ 186-
187-
188-
189-
190-
191 -
' -/
192-
• n—i—i—i—i—i—i—i—i—i—i—i—i—i—r~^
192 190 188 186 184 182 180 178
OH .
V/ppm
www.pdfgrip.com
302 Two-dimensional NMR spectroscopy
d1 P1 p2 p3
www.pdfgrip.com
Overhauser and magnetization transfer based spectroscopy 303
8 7 6 5 4 3
1
H/ppm
from the problem that correlations can occur through coupling, arising
from the same process as is exploited in TOCSY. Consequently more
caution is necessary in interpreting the results. The key observation
in Fig. 9.38 of the correlation between the CH group and the A:H
proving the relative correlation between these two protons, is safe as
no coupling pathway exists between them.
9.6.4 HOESY
Two-dimensional NOE can be used to link different species of nuclei,
when the technique is called HOESY (Heteronuclear Overhauser
Effect SpectroscopY). The technique works well whenever there is a
significant dipole-dipole relaxation process operating between the
nuclei. This normally means pairs of nuclei such as !H and 13C or 31P,
but in favourable cases quadrupolar nuclei can be involved. The
example chosen here is of 6Li-lH HOESY. Due to the low quadrupole
moment of 6Li, dipole-dipole relaxation is an important relaxation
pathway, and hence the Overhauser effect is observed. The compounds
www.pdfgrip.com
304 Two-dimensional NMR spectroscopy
(9.6) (9.7)
P3 P4
Figure 9.39 The HOESY pulse sequence, dl is for relaxation, d2 is the mixing
time for the NOE to develop, pi, p2, and p4 are 90° pulses, f/2 is the second
dimension for the two-dimensional experiment, and p3 is a 180° pulse.
www.pdfgrip.com
Inverse detection 305
JL
www.pdfgrip.com
306 Two-dimensional NMR spectroscopy
Table 9.1 The intensity gains on using inverse detection of several nuclei,
X, by 1H. Two gains are given. ("W/x)572 *s tne §am relative to simple
detection and (^/p/Yx)372 *s tne £am relative to an INEPT or DEPT spec-
trum.
Observe nucleus 31p
13 15 29 m 15
Inverse detect C N Si Rh N 103
Rh
(V7x) 5/2 31.6 306 56.8 5634 31.9 587
nuclei such as 13C have shifts spread over a wide frequency range. In
order to obtain reasonable resolution in the fj/jfj dimension, a large
number of spectra have to be acquired.
We will describe two inverse detection experiments, HMQC or
Heteronuclear Multiple Quantum Coherence and HMBC or Hetero-
nuclear Multiple Bond Correlation. HMQC is the simplest 1H,X corre-
lation experiment and is normally combined with BIRD selection (see
section 6.3.2) of the magnitude of the coupling to be observed and
GARP decoupling of the X nuclei (see section 6.1.5) and it is this
combination which is given here. The pulse sequence is given in
Fig. 9.41. Delay d2 is l/n/(XH) and d3 is set so the 1H spin vectors
are out of phase. HMQC is an alternative to heteronuclear COSY (see
section 9.3) but is generally more sensitive.
HMBC is used to study the two or three bond coupling correlations.
This can be done using HMQC or :H,X COSY (see section 9.3) but
HMBC suppresses more effectively the one bond coupling correlation.
The pulse sequence is given in Fig. 9.42. Both sequences start with the
BIRD sequence which, for HMQC is tuned to the coupling interaction
desired and for HMBC is tuned to ^(XH), the smaller couplings being
selected by d4.
Figure 9.41 The pulse sequence used for the HMQC experiment with BIRD selection and GARP decou-
pling, dl is a relaxation delay, pi, p3, p4, p8, and p9 are 90° pulses, p2, p5, and p6 are 180° pulses, d2 is
1/27, d3 is the BIRD delay, optimized for minimum 1VL NMR signal, and ^ is the variable delay for the
second dimension.
www.pdfgrip.com
Inverse detection 307
Figure 9.42 The pulse sequence used for the HMBC experiment with BIRD selection, dl is a relaxation
delay, pi, p4, p7, p8, and p9 are (90°)^ pulses, p3 is a (90°).,, pulse, p2, p5, and p6 are 180° pulses, d2 is
1/21/, d3 is the BIRD delay, optimized for minimum 1H NMR signal, d4 is l/2n/ where n = 2 or 3, d5 =
d2 - d4, and ^ is the variable delay for the second dimension.
(9.8)
1
H/ppm
Figure 9.43 A 500 MHz 1H-{3C HMQC NMR spectrum of [Ru(-q5-C5Me5)
(Ti5-2-C4H3S-2/,6/-C4H2S-2-C4H3S)], (9.8), in (CD3)2CO. The experiment
permits the assignment of the hydrogen bearing carbon atoms. (Reproduced
with permission from Graf et al (1995) Inorg. Chem., 34, 1562, copyright
(1995) American Chemical Society.)
www.pdfgrip.com
308 Two-dimensional NMR spectroscopy
80
90
100 _,
CO
110§
•O
•120
•130
140
7.5 7.0 6.5
1
H/ppm
-5850 £
Ol
-o
c*
T3
-5800 3
-5750
6 5 4 3 2 1
1
H/ppm
Figure 9.45 400 MHz lH,l95Pt two-dimensional inverse correlation NMR spec-
trum for the two isomers of [Pt(SnCl3)(Ti3-C4H7)(Ti2-PhCH=CH2)]. The normal
!
H NMR spectrum is superimposed on the spectrum. The 195Pt chemical shifts
are referenced to H2PtCl6. (Reproduced by permission of the American
Chemical Society, from Pregosin et al (1988) Organometallics, 7, 2130, copy-
right (1988).)
www.pdfgrip.com
Inverse detection 309
5(1H) = -16.28
•100
100
Hz
-16.2 -16.4
1
(a) H/ppm
8(31P) = 36.1 ;
1291
1292
CJl
1293 -^
(D^
1294 •§
1295
1296
www.pdfgrip.com
310 Two-dimensional NMR spectroscopy
X
p7 p8 p9 p10
Figure 9.47 The pulse sequence used for the HSQC experiment with BIRD selection, dl is a relaxation
delay, pi, p5, p8, and p9 are (90°)x pulses, p3 and p5 are (90°) pulses, p2, p4, p6, p7, and plO are 180°
pulses, d2 is 1/41/', and ^ is the variable delay for the second dimension.
www.pdfgrip.com
Three-dimensional NMR spectroscopy 311
3-
4-
H^~
1
Q.
,0.
I
" 5
6•* i i
10 9 . 8 7
(a) ^/ppm/^
4-
s-T
Q. .
^Q.
I
5-
6-
10 9 H1 8 7
(b) H/ppm//3
Figure 9.48 Two- and three-dimensional 600 MHz :H NMR spectra of inter-
leukin-lp. In both cases, NOESY is used to identify the connection 15
between
the NH and the aCH protons of individual amino acid residues. N editing
was used so that only 15NH groups are observed, (a) The two-dimensional
NOESY spectrum showing the large number of responses from the amino
acid residues, (b) A three-dimensional NOESY spectrum with the 15N chem-
ical shift being used as the /2 direction. One slice is plotted corresponding to
a 8(15N) = 123.7. The introduction of the third dimension has considerably
simplified the problem. (Reproduced with permission from Bax et al (1989)
Biochemistry, 28, 6150, copyright (1989) American Chemical Society.)
www.pdfgrip.com
312 Two-dimensional NMR spectroscopy
and 187Os = 1.15 x 10~3]. An example is shown in Fig. 9.46 where the
57
Fe signal of [Fe(Ti5-C5H5)H(PMe3)2] is observed through (a) 1H or
(b) 31P. The 1H spectrum of the 57Fe containing species is a triplet of
doublets though some intensity shows through from the non 57Fe
complex, which is almost 50 times more abundant, to give a triplet of
triplets. The 57Fe resonance is also a triplet of doublets. In (b) only
the couplings to P and H show.
9.7.3 HSQC
The HSQC (Heteronuclear Single Quantum Coherence) experiment
produces similar correlations to those obtained using the double
quantum experiment, HMQC, but has the advantage of removing WH
coupling leading to sharper signals. This offers considerable advan-
tages for larger molecules with crowded spectra. The pulse sequence
is given in Fig. 9.47.
9.9 QUESTIONS
www.pdfgrip.com
Questions 313
Bu* Bu*2
Assign the 119Sn signals and give the chemical shifts of each
different type of tin. Account for the multiplicity of the signals.
Derive the /(119Sn19F) values, where observed.
9.2. Figure 9.50 shows the 31P and 195Pt NMR spectra of a mixture of
two products formed by the addition of H~ to
www.pdfgrip.com
314 Two-dimensional NMR spectroscopy
(a)
• i i p i rrp-m-| i 1 1 1 , 1 1 . 1 p i i i | M i i p M i | i i i i p i i i p i i i p i i i p i i i p i i i | i i i i | M i i | i i i i p i i i p i i i p i i i | i i i i p M i pn . I , . . . . , ,
(b) JJL
\ l^ r
40 35 30 25 20 15 10 5
8(31P)
Figure 9.50 The NMR spectra of the two products of the reaction of [{Ti2-(Ph3P)2Pt}{7i6-Mo(CO)3}C7H5]+
with H-. (a) 64 MHz 195Pt NMR spectrum, referenced to K2PtCl4 in D2O. (b) 121 MHz 31P NMR spec-
trum. (Reproduced with permission from Jones et al (1996) Organometallics, 15, 596, copyright (1996)
American Chemical Society.)
9.3. Figure 9.51 shows the cyclopentadienyl region of the !H and 13C
NMR spectra of [{7i5-C5H4P(OC6H4NMe)2}Fe(CO)2Me]. Account
for the number of signals and derive the chemical shifts of the
signals and in the case of 13C the /(31P13C) values.
www.pdfgrip.com
Questions 315
1 1
(a)
1 ' ' ' 1 ' ' 1
1 i I > i i I i i
5.4 5.2 5.0 4.8 4.6
8(1H)
(b)
^lAri^*Vw/vVV^VW^Vv''/s'v^^^
I . . . . I >
97.5 95.0 92.5 90.0 87.5 85.0
13
5( C)
Figure 9.51 The cyclopentadienyl region in the (a) 300 MHz JH NMR spectrum and (b) the 75 MHz
} NMR spectrum of [{ri5-C5H4P(OC6H4NMe)2}Fe(CO)2Me] in CDC13. (Reproduced with permission
from Nakazawa et al. (1998) /. Am. Chem. Soc., 120, 6715, copyright (1998) American Chemical Society.)
p2_p3
www.pdfgrip.com
316 Two-dimensional NMR spectroscopy
9.5. Figure 9.53 shows the !H NMR spectrum of the SMe2 signal of
[Pt 2 (C 6 H 4 CH=N-l-C 6 H 10 -2-N=CHC 6 H 4 )Me 4 Br 2 (^-SMe 2 )],
complete with 195Pt satellites. Account for the coupling pattern.
www.pdfgrip.com
Questions 317
80 70 _. 60 50
8(31P)
(d)_
(c)_
(b)
(a)
-50 0 50 Hz
Figure 9.55 The magnetic field dependence of the *H NMR signal of
H4 of [HB(5-Bul-pyrazolyl)3Tl] in CDC13 at room temperature, (a)
200 MHz. (b) 300 MHz. (c) 400 MHz. (d) 500 MHz. (Reproduced with
permission from Parkin et al (1998) /. Am. Chem. Soc., 120, 10416,
copyright (1998) American Chemical Society.)
www.pdfgrip.com
318 Two-dimensional NMR spectroscopy
w_
(a)
www.pdfgrip.com
Questions 319
x
olL li il JlL kJL
110 100 90 80 70 60 50 40
31
6( P)
Figure 9.57 The 100.6 MHz ^P^H) NMR spectrum of [{(dppe)(2,6-Me2
C6H3NC)Pt)2Hg] in CD2C12. x marks impurities. (Reproduced with
permission from Puddephatt et al. (1996) Organometallics, 15, 1502,
copyright (1996) American Chemical Society.)
www.pdfgrip.com
320 Two-dimensional NMR spectroscopy
(e)
(d)'
(c)<
(b)~
2+
www.pdfgrip.com
Questions 321
Figure 9.59
2+
The partial 40.56 MHz 15N NMR spectrum of [Ru(NH3)4
(2-bzpy)] in d6-DMSO at room temperature recorded using the INEPT
pulse sequence. (Reproduced with permission from de Paula et al. (1999)
Polyhedron, 18, 2017, copyright (1999), with permission from Elsevier
Science.)
I T T
300 200 100 0 -100 -200 Hz
www.pdfgrip.com
322 Two-dimensional NMR spectroscopy
I I I I
0.5 0.4 0.3 0.2 0.1 0.0
8(1H)
Figure 9.61 The methyl region in the 400 MHz !H NMR spectrum of
[AsH(SiButMe2)2] in CDC13. The spectra are at (a) 10°C, (b) 20°C, (c)
30°C, (d) 40°C. (Reproduced with permission from Westerhausen et al
(1998) Inorg. Chem., 37, 619, copyright (1998) American Chemical
Society.)
www.pdfgrip.com
Questions 323
(j)
(f)
(e)
(0_ (d).
(c)
(b)
www.pdfgrip.com
324 Two-dimensional NMR spectroscopy
9.14. Figure 9.63 shows the central signals of the variable temperature
29
SipH} NMR spectra of ds-[Pt(PEt3)2(SiFMe2)2]. The 195Pt satel-
lites are outside the spectral range. Explain the multiplicity of
the spectra at -80°C and 40°C. Account for the changes in the
spectra with temperature. Suggest a mechanism of fluxionality
which is consistent with the spectra.
I I
64 62 60 58
5(29SJ)
Figure 9.63 The 79.3 MHz 29Si NMR spectrum of ds-[Pt(PEt3)2(SiFMe2)2]
in d8-toluene at (a) -80°C, (b) -50°C, (c) -30°C, (d) 20°C, (e) 40°C
(Reproduced with permission from Tsuji et al. (1998) Organometallics, 17,
507, copyright (1998) American Chemical Society.)
www.pdfgrip.com
Questions 325
(a)
(b)
7.2 7.1 7.0 6.9 6.8 6.7 6.6 6.5
8(1H)
www.pdfgrip.com
326 Two-dimensional NMR spectroscopy
^--30
|--20
|--10
I- o
I- 10
i- 20
Figure 9.66 The 80.2 MHz 11E{1H] COSY NMR spectrum of [(9-BBN)
B10H12]~ in ^7-DMF. (Reproduced with permission from Shore et al.
(1998) Inorg. Chem., 37, 3276, copyright (1998) American Chemical
Society.)
www.pdfgrip.com
Questions 327
www.pdfgrip.com
328 Two-dimensional NMR spectroscopy
A/1
(b)
(a)
www.pdfgrip.com
Questions 329
(en
[!_
(b)^_
(*!>
www.pdfgrip.com
330 Two-dimensional NMR spectroscopy
Prj3P l/PPr's
NCMe
(a)
I
-20 -21 -22 -23 -24
5(1H)
24
22 5(1H)
20
-20 -22 -24 -20 -22 -24
5(1H) 6(1H)
Figure 9.71 (a) The 300 MHz 'H NMR spectrum of the hydride signals of
[Ir2H4(N2C3H3)2(NCMe)(PPri3)2]. (b) The COSY spectrum, (c) The NOESY
spectrum. (Reproduced with permission from Oro et al. (1998)
Organometallics, 17, 683, copyright (1998) American Chemical Society.)
www.pdfgrip.com
Questions 331
353 K
343 K
333 K
323 K
iaig
293 K
283 K
273 K
CH6^
263 K
253 K
243 K
233 K
NH -
223 K
(a) 213 K
203 K
;193K
www.pdfgrip.com
332 Two-dimensional NMR spectroscopy
X 1
I | I I I I ' ' ' 'I
0 -5 -10
(d) 8(1H)
www.pdfgrip.com
Questions 333
www.pdfgrip.com
www.pdfgrip.com
Magnetic resonance
imaging and
biomedical NMR
10
Magnetic resonance imaging has made immense strides over the last
decade. Initially, it was applied to medicine where it has proved to
give an extremely useful extra diagnostic method to radiographers and
has now become in addition, a medical research tool of increasing
usefulness. If these were its only uses, imaging might not be an appro-
priate subject for a book aimed at chemists but applications in biology
and technology are now well established and a basic understanding at
least of the subject is necessary in today's world for the aspiring NMR
spectroscopist. Note that the word 'nuclear' has been quietly dropped
from the description of the technique. This is because the public has
difficulty in understanding the difference between stable and unstable
isotopes. It also avoids any confusion with Nuclear Medicine, which
discipline does use unstable isotopes. We will describe briefly how
images are obtained, remembering that there now exist a multitude of
RF pulse/magnetic field gradient sequences designed to obtain various
ends, and then look at a number of more chemically oriented appli-
cations. Biomedical NMR looks at the chemistry taking place in living
matter and, apart from straightforward spectroscopy, may use imaging
with spatially resolved spectra or simply place a coil on the surface of
a sample near an organ of interest, and watch what happens when
various constraints are imposed on the system.
www.pdfgrip.com
336 Magnetic resonance imaging and biomedical NMR
information for us. Since the human body contains a large proportion
of water, it became evident that, if one could map its spatial distrib-
ution, then one might be able to investigate the nature of the soft
body tissue. Water is present as about 55% of body weight and the
proportion varies widely in different parts of the body, as does
the water mobility, and so its relaxation rates. In general, in body fluids
T2<T^ We will describe first a basic experiment to indicate the general
principles used in obtaining an image following Fig. 10.1.
The basic idea behind producing an image can be understood as
follows. An object is three-dimensional but the display of its image
has to be two-dimensional. It is necessary therefore to first define a
plane whose two-dimensional image is to be reproduced. This is done
by subjecting the object to be imaged to a field gradient (and so
frequency gradient) in one direction only, and we will take the z direc-
tion, which may be vertical or horizontal in the case of imaging. A 90°
pulse is applied, but with a fairly long duration or with a gaussian
shape, to give a small frequency spread, so that the nuclei are in
resonance and swing into the xy plane over only a short distance
along the magnetic field gradient. For instance, if the field gradient
Magnetic
field
gradient in |z| Directions
Slice
selective_
pulse
Figure 10.1 Pulse and field gradient timing for obtaining a proton density
image of a slice of a subject. The finishing point of the signal produced for
each volume element, or voxel, under the x gradient depends on the x coor-
dinate of the resonating fluid and this is encoded into the signal produced
under the y gradient as a phase change. The output thus contains informa-
tion of both x and y coordinates of every part of the subject in the slice
defined by the z gradient and RF pulse
www.pdfgrip.com
Whole body imaging 337
dB0/dz = 5 juiT mm'1 along the main field axis, then the frequency
change for 1H that occurs for a 2 cm displacement along the z axis is
4150 Hz, and a rectangular pulse of the order of 240 JJLS long will define
a slice of the object 2 cm thick. The z field gradient is switched off at
the end of the pulse and, say, gradient x is then switched on. The
water in the 2 cm slice thus produces a signal, the frequency of which
depends upon the x coordinate only so that the rows of spin density
perpendicular to the gradient direction each have different frequency.
The magnetization evolves with time and, after some time tx9 that of
the row at each x coordinate will have a particular phase, which is a
function of tx and x coordinate. This is often called the phase encoding
gradient for this reason. The x gradient is then switched off and
replaced by a y gradient. The output frequencies now depend upon
the y coordinate but with phase determined by the x coordinate. The
output is collected in the time domain t. Outputs are collected for
several values of tx and the data are then subjected to a two-dimen-
sional Fourier transform, to give a map of proton density in the slice
originally defined by the combination of RF pulse and z gradient. The
timing of the experiment is summarized in Fig. 10.1. In practice, the
technique shown has a number of drawbacks, principally because
the relaxation times T2* of body fluids are rather short due to the non-
homogeneous nature of the subject so that the signal has very little
intensity by the end of the collection period and the intensity is also
a function of tx. The example given is however, clear and follows
directly from the techniques for two-dimensional spectra already
discussed.
Many different sequences have now been developed for whole body
imaging which avoid the disadvantages mentioned above and also
permit contrast weighting of the image by utilizing the differences in
7\ or T2 of body tissue. In addition the effects of blood flow and the
unavoidable motion of body tissue have to be compensated for and
this necessitates adding further complexity to the sequences used. We
will describe one basic sequence here which is shown in Fig. 10.2. The
first point to note is that the slice selective z gradient (Gz) is followed
by a period of z gradient of opposite sign. During the short slice selec-
tive period, the spins lose some phase coherence in the xy plane
through the depth of the slice and this reversal of gradient compen-
sates for this, refocusing the spins. Secondly, tx is kept constant and
the changes in phase encoding are brought about by varying the inten-
sity of the phase encoding x gradient (GJ, which is shown as a pulse
cut by horizontal lines. The y gradient (Gy) is next switched on to give
the frequency encoding, though by the end of this time the signal will
be of almost zero intensity. A 180° RF pulse is then applied, in conjunc-
tion with a z gradient so that it is also slice selective and this produces
www.pdfgrip.com
338 Magnetic resonance imaging and biomedical NMR
90° 180°
Echo
RF
Slice selection
Gz
Phase encoding
Gx
Frequency
encoding
Gy
time
Figure 10.2 A typical sequence used for whole body imaging. The vertical
arrow implies that the various phase encoding gradients are applied sequen-
tially. Gz, Gx, Gy, are the magnetic field gradients applied along the three
axes of the system.
www.pdfgrip.com
Whole body imaging 339
Figure 10.3 Two sections of a human head showing the location of a brain
tumour (Images supplied by Bruker Medizintechnik.)
www.pdfgrip.com
340 Magnetic resonance imaging and biomedical NMR
www.pdfgrip.com
Diffusion and flow 341
of the brain, whereas the opposite was the case when faces were in
his field of view. The centres of activity are quite dispersed and some
of the centres are involved in more than one activity.
90° 180°
g g
RF Echo
Figure 10.5 The basic pulse sequence used to measure the spatial displace-
ment of spins in short intervals of time. The two RF pulses of 90° and 180°
are the usual Carr-Purcell sequence and two short pulses of magnetic field
gradient (g) are added equally displaced from the 180° RF pulse. This
sequence is followed by an echo whose intensity is reduced if there is displace-
ment of spins between the two pulses g.
www.pdfgrip.com
342 Magnetic resonance imaging and biomedical NMR
faster in the centre of the conduit than at its walls. It will be evident
by referring to Fig. 10.2 that this pulse sequence can easily be combined
with an imaging sequence and so measure flow or diffusion through
an image. Other methods are also possible and are often called time
of flight measurements. Thus if the repetition time between pulse
sequences is short then stationary spins are saturated and flowing
spins are not and so appear with different intensity. Alternatively, if
all the spins in a slice are inverted by a 180° pulse and then a waiting
period given such that all the stationary spins are at null intensity, new
spins flow into the slice wherever there is flow and these appear in
the image.
Diffusion weighted images of the brain have been found to be invalu-
able in the early diagnosis of stroke damage since the diffusion of
water is slower in ischemic tissue and, indeed, imaging is the only tech-
nique capable of doing this. Other medical applications include the
measurement of blood flow in phase contrast magnetic resonance
angiography. If flow occurs in a linear gradient there is a phase change
induced which is identical for every position along the gradient and
allows the flowing fluid to be differentiated from stationary fluid. The
magnitude of the phase change depends upon the timing of the pulse
sequence, the value of the gradient, and the velocity, which can thus
be calculated from the data.
www.pdfgrip.com
Chemical shift imaging 343
RF FID
Figure 10.6 The WATERGATE pulse sequence. The two tall RF pulses are
non-selective and the two broad, short ones are selective and at the water
frequency. The two magnetic field gradient pulses G refocus only the non-
water resonances.
www.pdfgrip.com
344 Magnetic resonance imaging and biomedical NMR
One can do even better than this, however, and produce the spectra
of all voxels in a slice using an imaging scan. The slice selective 90° -
T - 180° sequence is used similar to that described above (Fig. 10.2)
but gradient pulses are applied along the other two axes, say x and y,
just before the 180° RF pulse, with many values of Gx and many values
of Gy for each value of Gx. Fourier transformation of this data stack
gives a spectrum for each voxel of the slice. The sequence has to be
preceded by some sort of water suppression sequence such as that
included in Fig. 10.8. Figure 10.7 shows such an image. It comprises
a matrix of 16x11 spectra at relatively low resolution but within
which it is possible to discern major changes over the space occupied
by the sample. The sample is the brain of an anaesthetized rat in
which a tumor had been implanted. Each spectrum corresponds to a
voxel dimension of 1 x 1 mm. The resonances which are visible are,
starting from low frequency, lactate in the tumour, 7V-acetyl aspartate,
Figure 10.7 1H spectra of a slice of a rat brain into which a tumour had
been implanted. Each spectrum represents a pixel of 1 x 1 mm. The tumour,
which is rich in choline and lactate, is located at the centre left of the figure.
Note that the oval shaped brain produces no responses in the lower right
and left corners of the matrix. (Figure provided by A. Ziegler, INSERM,
Grenoble. To be published in Magn. Res. in Medicine)
www.pdfgrip.com
Chemical shift imaging 345
Figure 10.8 X H NMR COSY imaging pulse sequence used to obtain the two-
dimensional spectra of Fig. 10.9. First the water is subjected to three selec-
tive pulses followed by pulses of gradient which dephase its transverse
magnetization, so that the water signal is much reduced; this is called the
CHESS sequence. The RF sequence that follows is a simple COSY sequence
but the magnetic field gradient pulses which are added produce an image.
The first 90° selective RF pulse is thus converted into a slice selective pulse,
and the second is accompanied by gradients in the three axes, several values
of Gx being used for each value of Gz. Note that the gradient pulses can
overlap.
www.pdfgrip.com
346 Magnetic resonance imaging and biomedical NMR
Figure 10.9 Correlation peak imaging of rat brain in which a tumour had been implanted, (a) Normal
brain tissue, (b) Tumour tissue. The cross peaks correspond to (1) Af-acetyl aspartate, (2) glutamine/gluta-
mate, (3) glucose, (4) aspartate, (5) taurine, (6) inositol, perhaps choline, (7) lactate, (8) alanine and (9)
hypotaurine. The nominal voxel volume for each spectrum is 23 fjil and slice thickness 5 mm. (Figure
provided by A. Ziegler, INSERM, Grenoble.)
each molecule and which are far better resolved than in the simple
spectrum. It is thus possible to detect some fourteen compounds in
the rat brain, spatially localized (Fig. 10.10). The change in the metabo-
lites present in tumour and normal tissue is very marked. Note that
the strong TV-methyl resonance of choline is not seen in these COSY
plots since there is no coupling. The choline resonance remains in the
diagonal.
www.pdfgrip.com
Biological uses of imaging 347
Figure 10.10 Metabolic images for a whole rat brain corresponding to the cross peaks of Fig. 10.9. The
contours of brain and tumour are superimposed on the images, (a) shows the distribution of Af-acetyl
aspartate, (b) that of glutamine/glutamate, (c) that of glucose, (d) that of lactate, (e) that of hypotaurine,
(f) that of alanine, (g) that of choline/inositol, (h) that of aspartate and (i) that of phosphoethanolamine.
(Figure provided by A. Ziegler, INSERM, Grenoble. To be published in Magn. Res. in Medicine.)
www.pdfgrip.com
348 Magnetic resonance imaging and biomedical NMR
Figure 10.11 Image of a cross-section of the hypocotyl of a castor bean seedling. The eight vascular bundles
(b) are connected by the meristem ring. The cellular structures of pith parenchyma (a) and cortex
parenchyma (c) are visible. (From Ziegler et al (1996) /. Magn. Reson., 112B, 141, with permission.)
www.pdfgrip.com
Industrial uses of imaging techniques 349
such things as the distribution of fluid flow in packed bed columns can
be visualized or the mixing of fluids of different NMR relaxation times
which have contrast in the images.
Some specific examples are the study of packed bed reactors. The
fluids present may be gas and liquid when it is easy to determine how
the two are distributed because of the different spin densities, and the
way the distribution changes with flow rate is easy to follow. Such
reactors are most efficient in the trickle flow regime when the liquid
falls in rivulets from one catalyst particle to the next. The rates of
reaction also depend upon the shape and size of the catalyst particles
and the way they pack in the reactor, and these factors can now be
studied directly. Packed reactors through which two-phase liquid
systems are being passed can also be examined and the way the phases
flow and co-dissolve can be observed. The distribution of flow rates
in the interstices of the reactor can be measured using techniques
already discussed. An example is given in Fig. 10.12.
The deactivation of catalyst beads can also be followed by imaging.
A catalyst bead, whose pores may have become partially blocked by
8.3 mm s~1
0.0 mm s~1
Figure 10.12 Flow rates of a water/hydrocarbon mixture through the beads (black) in a packed bed reactor.
The rates are calibrated by the shade scale at the left. (From Gladden (1998) Chem. in Britain, 34, No.
3, 35, with permission.)
www.pdfgrip.com
350 Magnetic resonance imaging and biomedical NMR
www.pdfgrip.com
Biomedical NMR 351
(a)
(b)
Figure 10.13 The 31P topical NMR spectra obtained at 32.5 MHz using a
single-turn coil placed on the surface of a subject's arm, in a contoured
magnetic field, (a) The normal spectrum (64 2-s scans). Peak assignments are
given in the text, (b) The same subject 50 min after the application of a tourni-
quet to the upper arm. (Reproduced with permission from Gordon (1981)
Eur. Spectrosc. News, 38, 21.)
www.pdfgrip.com
352 Magnetic resonance imaging and biomedical NMR
[1,2-13C2]glycerolC1+C3
(B)
[1-13C]glycerolC1+C3
(A)
Vw^^^^X^' V*vyR\*vX^rv^*w~»*
ppm 100 80 60 40
Figure 10.14 In vivo 13difference 13C spectra of13rat liver after 125 minutes of
infusion of either [1- C] glucose (A) or [1,2- C] glucose (B). The doublets
indicate C-C bonds which have been metabolized without bond fission.
Resonances 1, 4 and 6 arise from glycogen, the others are named
in the text. (From Kiistermann et al (1996) Bruker Report 143, 33, with
permission.)
www.pdfgrip.com
Biomedical NMR 353
Echo
RF-
Gradient slice 1
Gradient slice 2
Gradient slice 3
Figure 10.15 The PRESS sequence for obtaining the spectrum of a single
voxel, modified to diffusion weight the spectra so obtained. The PRESS pulses
are open, the diffusion weighting pulses are marked with a diagonal. D is the
interval between pairs of these pulses and d is the pulse duration.
www.pdfgrip.com
354 Magnetic resonance imaging and biomedical NMR
www.pdfgrip.com
High-resolution
solid-state NMR 11
We have already mentioned in Chapter 4 that, in the solid state,
the relaxation time 7\ is long due to the lack of modulation of the
dipole-dipole interaction and T2 is short due to mutual spin flips occur-
ring between pairs of spins. In a static solid, each nucleus produces a
rotating magnetic field as it precesses in the applied magnetic field,
and this can cause direct exchange of energy between nuclei. The life-
times of the spin states are thus reduced and so T2. In addition, each
spin has a static field component that influences the Larmor frequen-
cies of its neighbours. An individual nucleus will experience the fields
of several neighbours, but their spin directions will vary randomly,
so that there will be a range of frequencies that will add to the line
broadening due to the rapid rate of relaxation. Finally, particularly for
the heavier nuclei, including 13C, there will exist a chemical shift
anisotropy, which will also contribute to the broadening, assuming that
the sample is a powder or a glass and not a single crystal, because the
chemical shift varies with orientation relative to the BQ direction. Thus
solid materials, particularly if they contain nuclei with high magnetic
moments such as *H or 19F, will have broad, structureless resonances,
which will not permit the type of investigation that we have shown
can be carried out in the liquid phase. This state of affairs has proved
a challenge to the NMR community, who have over the last two
and a half decades found means to render ineffective the apparent
physical restraints to the spectroscopic examination of solids at high
resolution.
Before discussing this work in detail, however, it is necessary to
mention two useful aspects of the broad lines. In the first place, because
the broadening is determined by the dipole-dipole interactions, it is
sensitive to the distance separating interacting spins. The spectrum of
a solid can thus be used to obtain internuclear distances, which in the
case of the proton are difficult to obtain by other means. The mole-
cules studied must be static in the solid state and must be sufficiently
simple, preferably an isolated spin pair, that the resonance width can
be interpreted in terms of a single, principal distance. If the molecules
reorient in the solid in some way, then this modulates the internuclear
interaction and its magnitude is reduced, and this is the second prop-
erty that proves useful. If the linewidth of a solid material is measured
as a function of temperature from very low temperatures, it is often
www.pdfgrip.com
356 High-resolution solid-state NMR
found that there are quite rapid changes in linewidth at certain tran-
sition temperatures. These mark, if the temperature is being gradually
increased, the onset of motion within the solid lattice. Figure 11.1
shows an example for the solid complex adduct of boron trifluoride
with trimethylamine, Me3N—>BF3. The !H resonance linewidth below
80K is 85 kHz (the old unit of 1 gauss is equivalent to 4250 Hz in this
case). Heating from 68 to 103K reduces the linewidth to about 21 kHz,
and this corresponds to the onset of rotation of the methyl groups
around the C-N bonds. Further narrowing to 13 kHz occurs on raising
the temperature to 150K owing to the onset of rotation of the whole
NMe3 moiety around the B^N bond. Finally, just below 400K, the
line narrows to a few hundred hertz as the whole molecule starts to
rotate and diffuse isotropically within the still solid crystal. The 19F
resonance can also be examined and is found to be broad only below
77K and the BF3 rotates around the B«-N axis at all higher temper-
atures. Of course, when the sample melts, the linewidth falls to a frac-
tion of a Hertz.
Such studies using broad lines are, however, of relatively limited
application, and solid-state NMR formed only a small part of the total
80 200 400
Temperature (K)
Figure 11.1 The proton resonance linewidth of the methylamine protons in the solid adduct (CH3)3N —> BF3
as a function of temperature. Note that 1 gauss = 10"4 tesla = 4250 Hz. (After Dunnell (1969) Trans.
Faraday Soc., 65, 1153, with permission.)
www.pdfgrip.com
Magic angle spinning 357
NMR work that was undertaken. The situation has now changed
dramatically with the application of the modern techniques to be
described below. Both I = 1/2 and quadrupolar nuclei are studied,
though the treatment of these two classes of nuclei is rather different,
and we will have to discuss them separately. One technique is common
to both, one which has revolutionized solid-state NMR more than any
other, and we will describe this first.
Bz = - L (3cos2e - 1)
www.pdfgrip.com
358 High-resolution solid-state NMR
IX
Figure 11.2 A line of the magnetic field originating from a magnetic dipole has zero z component at a
point situated on a line originating at the centre of the dipole and at an angle of 54°44' to the direction
of the dipole. (From Bruker CXP Application Notes, with permission.)
Figure 11.3 Showing how a solid sample is mounted for magic angle spinning and how this gives the inter-
nuclear vectors an average orientation at the spinning angle. (From Bruker CXP Application Notes, with
permission.)
www.pdfgrip.com
Spin-1/2 nuclei with low magnetogyric ratios 359
(a)
(b) (c)
Figure 11.4 The 13C spectra of solid adamantane (formula inset), (a) With MAS. (b) With high-power
proton double irradiation, (c) With both MAS and double irradiation. (From Bruker CXP Application
Notes, with permission.)
www.pdfgrip.com
360 High-resolution solid-state NMR
in this way has certain drawbacks. The rather long 13C Tl values mean
that pulse rates have to be sufficiently slow so as not to saturate the
resonances, and the natural insensitivity of the nucleus means that long
accumulation times are needed. These disadvantages can be circum-
vented by a modification of the double-resonance technique, which
permits the exchange of polarization between the !H and 13C spins,
and which is called cross-polarization (CP); when combined with MAS,
the whole is abbreviated to CP-MAS.
In order that energy exchange shall be possible between the two
nuclear species, we must introduce components of motion with the
same frequency for each. This is done as follows. Referring to
Fig. 11.5, we first prepare the protons for the cross-polarization by
applying a short 90° pulse, which swings the proton magnetization into
the xy plane. We will call this field Bm. We then change the phase of
B1H by 90° (and perhaps reduce its intensity also), so that the BIH
vector becomes parallel with the magnetization in the xy plane, which
therefore remains in the xy plane and stays there locked to Bm. This
second pulse is known as a spin locking pulse. The magnetization
precesses in the xy plane at the Larmor frequency, and can be thought
of as also precessing around BIH at a frequency of yHBm/2TT Hz,
behaving as if it was very strongly polarized in the weak BIH. Reference
to equation (1.3) will show that such a polarization, correct under
normal circumstances for Z?0, can only be attained if the temperature
is very low for a weak field BIH, and we can consider that the spin
locking has cooled the spins, which can now act as an energy sink.
Energy transfer is obtained by applying a long pulse at the 13C
frequency, U1C, which has an amplitude such that the 13C nuclei precess
around it at a frequency of ycBlc!2^ Hz which is equal to yHBlH/2^
Hz. This is known as the Hartmann Hahn matching condition. The
identical frequency components allow energy transfer, which follows
an exponential curve, and when this has reached a maximum the BIC
is cut off and is followed by a 13C FID of enhanced intensity. The i?1H
field remains on during this time and provides the decoupling of the
protons from the carbon nuclei. The experiment can be repeated when
the proton spins have reached equilibrium again and so the effective
relaxation time is the shorter one of the proton system. The cross-
polarization increases the 13C population difference by the factor
'Yi/Yc and so produces a useful improvement in signal strength. An
example of a 13C CP-MAS spectrum is shown in Fig. 11.6 for a more
complex molecule, the steroid deoxycholic acid, together with a partial
assignment of the resonances. Deoxycholic acid forms inclusion
compounds, and, when the guest molecule is ferrocene, [Fe(T]5-C5H5)2],
the methyl singlets due to CIS, C19 and C21 become doublets due to
differentiation of the deoxycholic acid molecules in the solid lattice.
The cross-polarization technique is much used with low--y nuclei where
the compound studied contains hydrogen; 13C, 15N, 29Si, 31P and 113Cd
are some examples of / = 1/2 nuclei studied. Quadrupolar nuclei can
also benefit from the CP technique.
www.pdfgrip.com
Spin-1/2 nuclei with low magnetogyric ratios 361
If the spinning speed is appreciably lower than the width of the static
resonance of the compound studied, then sidebands are produced sepa-
rated by the spinning speed. Figure 11.7 shows the solid-state 31P
spectra of aminomethanephosphonic acid, H2NCH2PO3tl2, which has
a zwitterionic structure. All the three spectra shown benefited from
Spin
Decoupling
lock
Irradiation
at 13 C
Frequency
1H
www.pdfgrip.com
362 High-resolution solid-state NMR
C10
C13
Figure 11.6 The 13C CP-MAS spectrum of deoxycholic acid. The 13C frequency was 50.32 MHz and that
for *H was 200 MHz. Some 350 mg of sample was packed into the MAS rotor and some 800 transients
were acquired. The contact time was 1 ms and recycle delay 3.5 s. (From Heys and Dobson (1990) Magn.
Reson. Chem., 28, S37-46, copyright (1990) John Wiley and Sons Ltd, reprinted with permission.)
CP, but the uppermost one (a) was from a static sample and shows
the shielding anisotropy of the phosphorus nucleus, which does not
have axial symmetry and so has three principal components crn, cr22
and a33 (see Chapter 2). MAS at 813 Hz (b) produces a group of
narrow lines separated by the spinning frequency. The number of lines
is reduced on increasing the spinning speed to 2950 Hz (c), and it is
evident that only one does not move, so that this (arrowed) is the
true resonance with the isotropic chemical shift and the remainder
are spinning sidebands. It is particularly important to note that at
low spinning speeds, the envelope enclosing the sidebands approxi-
mates the static lineshape and so retains the form of the shielding or
chemical shift anisotropy. The 31P spectrum of Af,Af-dimethylamino-
methanediphosphonic acid, Me2NHC(PO3H2)2H, is also shown for a
spinning speed of 1740 Hz (d). In this case, there are two resonances
that are not sidebands, since the two phosphorus atoms in a single
molecule are not related by symmetry due to the crystal structure. In
solution, of course, only a singlet 31P resonance is produced by this
compound.
www.pdfgrip.com
Spin-1/2 nuclei with low magnetogyric ratios 363
(a)
(b)-
(c)
www.pdfgrip.com
364 High-resolution solid-state NMR
example is shown in Fig. 11.8 for both a plastic and a coal and, while
the resolution for the coal is not exceptional, it has to be remembered
that it is a most complex mixture of structures and that it is very useful
to be able to distinguish aromatic and aliphatic resonances and perhaps
to be able to do this quantitatively.
It will be no surprise to find that two-dimensional techniques have
also been found useful for the study of solid state systems. We give
an example of an EXSY experiment used to study motion in the
solid polymer [(Me3Sn)4Ru(CN)6]00; in which the tin coordination is
(a)
(b)
\ '
250 200 150 100 50 -50
8 (ppm)
Figure 11.8 The 13C CP-MAS spectra at 15 MHz of (a) a sub-bituminous coal
and (b) polycarbonate solids. (From Wind (1991) Modern NMR Techniques
and Their Application in Chemistry, eds Popov and Hallenga, Marcel Dekker
Inc., New York, p. 186, with permission.)
www.pdfgrip.com
/ = 1/2 nuclei with high magnetogyric ratios 365
F2 / PPM
-4-
-3-
-2 -
-1 -
0-
1-
2•
3-
4-
5-
6-
7-
8-
9-
10-
1 0 9 8 7 6 5 4 3 2 1 0 - 1 - 2 -3 -4
F1 / PPM
www.pdfgrip.com
366 High-resolution solid-state NMR
are the static linewidths very large, but the chemical shifts are small,
so making big demands on the resolution ability of the system. MAS
at the very highest spinning speeds and using the highest magnetic
fields to maximize the chemical shift dispersion is capable of reducing
a static linewidth of the order of 10 kHz to around 1500 Hz, and
this can give resolvable resonances though the actual improvement
obtained is very sample dependent. Alternatively, the spins can be
swung around by a succession of pulses so that they appear to adopt
the magic angle in the rotating frame. A typical sequence, called
MREV-8, is shown in Fig. 11.10 and has the effect that the magneti-
zation is shifted quickly between the three orthogonal axes, which, as
it will be remembered from section 4.3 and Fig. 4.7, are placed at the
magic angle from their three-fold symmetry axis. The spins thus hop
around the magic angle axis and their dipole-dipole interaction is
much reduced, though the chemical shift anisotropy and heteronuclear
(I)
(II)
20 10 0 8(ppm)
www.pdfgrip.com
/ = 1/2 nuclei with high magnetogyric ratios 367
OH)
(f) Expansion
10
(e) CRAMPS
(a) Static
40 20 0 -20 -40
kHz
Figure 11.10 (I). The eight-pulse MREV-8 cycle. Each pulse is a 90° pulse,
which rotates the magnetization around the x or y axes in one direction or
its opposite. The large spacings are double the length of the small spacings.
The nuclear signal is sampled between pulses. (II) The *H CRAMPS spec-
trum at 200 MHz of aspartic acid HOOCCHNH 2CH2COOH. (From Bruker
Report 1/1988, with permission.) (Ill) ]H NMR spectra of solid adipic acid
(a) static, showing the broadening due to dipolar interactions, (b, c, d) with
MAS at various speeds, none exceeding the linewidth of the static spectrum
(40 kHz) and (e) with CRAMPS, (f) is an expansion of (e) and it will be seen
that the three types of proton are well resolved. (From Maciel, in Gerstein
(1996) Encyclopedia of Nuclear Magnetic Resonance, 3, 1501, eds Grant and
Harris, copyright John Wiley and Sons Ltd, with permission.)
www.pdfgrip.com
368 High-resolution solid-state NMR
Quadrupolar nuclei in the solid state usually have weak dipolar inter-
actions with their surroundings, and this will not concern us here.
Where they do exist then a cross-polarization experiment may be
www.pdfgrip.com
MAS of quadrupole nuclei 369
142.8
32.0
14.2
• 5.4
• 2.1
0.94
0.18
0.0015
-150 -160 -170 -180 -190 -200 -150 -160 -170 -180 -190 -200
8F/ppm
www.pdfgrip.com
370 High-resolution solid-state NMR
the energy of one transition and increase the energy of the other. The
degeneracy of the three possible transitions is now lifted and we detect
three resonances. This is shown schematically in Fig. 11.12. If we
change the orientation of the crystal in the magnetic field, we change
the orientation of the EFG tensor and so the interaction of the nucleus
with the EFG. This alters the energy levels: though the l/2<-»-l/2 tran-
sition frequency is unchanged, the other two will both be affected, one
being increased and one decreased. At certain orientations (say 0°)
they will coincide with the 1/2^-1/2 transition and, as the crystal is
rotated, they will move away from this l/2<-»-l/2 central line, reach a
maximum displacement, return to the central line, cross over, reach
maximum again and, at 180° rotation for crystals with axial symmetry,
again coincide with the central line. This is known as the first-order
Figure 11.12 Energy level diagram for a nucleus / = 3/2. The transition ener-
gies are equal if only the magnetic field BQ is taken into account, but the
quadrupole interaction causes one to become larger, one to become smaller
and one, the central transition, to remain unchanged.
www.pdfgrip.com
MAS of quadrupole nuclei 371
200-
-200
60 120 180
Degrees rotation
Figure 11.13 The positions of the two satellite lines in the 23Na spectrum of
a single crystal of NaNO3 as a function of angle between the B0 magnetic
field direction and the three-fold symmetry axis of the crystal. The origin of
the frequency axis is the frequency of the central line. (Derived from Andrew
(1958) Nuclear Magnetic Resonance, Cambridge University Press, Cambridge.)
www.pdfgrip.com
372 High-resolution solid-state NMR
+
v2 = - (^[W !) - IJC1 - cos26)(9cos2e - 1)
www.pdfgrip.com
MAS of quadrupole nuclei 373
0 ± JL ±
2 1 6 3
www.pdfgrip.com
374 High-resolution solid-state NMR
(a) (d)
(b) (e)
(c) (f)
ppmfrom Al (H2O)g3+
Figure 11.15 The 27A1 solid-state spectra of the tridecameric aluminium cation
at different magnetic fields, with and without sample spinning. Spectrometer
frequencies were (c), (f) 39 MHz, (b), (e) 93.7 MHz and (a), (d) 129.7 MHz.
Samples (e) and (f) were static, samples (a), (b) and (c) were spinning at the
magic angle, and sample (d) was spinning at an angle of 75° to the magnetic
field direction. (After Oldfield et al (1984) J. Magn. Reson., 60, 467, with
permission.)
www.pdfgrip.com
MAS of quadrupole nuclei 375
80 70 60 50
8 (ppm)
Figure 11.16 The 27A1 MAS spectra at 130 MHz of CaO3Al2O3-3H2O showing
the two superimposed signals and their computed line shapes. Note particu-
larly the two close values of the true isotropic chemical shifts, which are
marked on the shift axis and correspond with hardly any signal intensity. Note
also how two types of chemically very similar aluminium are well differenti-
ated by different quadrupole couple constants. (Reprinted with permission
from Muller et al (1986) /. Chem. Soc., Dalton Trans., 1277.)
www.pdfgrip.com
376 High-resolution solid-state NMR
Figure 11.17 The 23Na MAS NMR spectra of the homonuclear alkalide
Na[C222.Na]. (From Dye et al (1991) Modem NMR Techniques and
Their Application in Chemistry, eds Popov and Hallenga, Marcel Dekker Inc.,
New York, p. 291, with permission.)
www.pdfgrip.com
Some applications 377
11.5.2 Zeolites
These substances are formed of networks of aluminosilicates that
contain pores of certain fixed sizes and are active as catalysts. They are
produced by crystallization from a gel formed upon mixing, say, an
aluminate salt with a soluble silicate, and the structure of the zeolite
formed depends upon the nature of the components used to form the
gel. The ratio of silicon to aluminium present in the zeolite can be
varied within certain limits and the silicon is always in excess. The solids
thus contain Si-O-Si and Si-O-Al linkages but not A1-O-A1 linkages,
which appear to be forbidden. These substances contain two magneti-
cally active nuclei, 29Si with / = 1/2 and the quadrupolar 27A1 with
I = 5/2, and both are used extensively in their study. The 29Si spectra
may contain up to five resonances, which correspond to tetrahedral
SiO4 units with zero, one, two, three or four attached aluminium atoms.
A spectrum of zeolite Na-Y with Si/Al = 2.61 is shown in Fig. 11.19,
where the resolution of the different environments is seen to be
www.pdfgrip.com
378 High-resolution solid-state NMR
10 15 20 25 3 5
Hydration time (min)
Figure 11.18 Hydration of calcium aluminate. A typical 27A1 MAS spectrum is shown at the top left of
the figure. The three other plots show the progress of hydration with time at three different temperatures.
The full curves show the heat evolution and the broken curves show the percentage of six-coordinate
aluminium formed. SB on the spectrum signifies spinning sidebands. (Reprinted with permission from
Rettel et al (1985) Br. Ceram. Trans. J. 84, 25.)
excellent. With care, the intensities are quantitative and the pattern
allows the Si/Al ratio to be calculated. In natural zeolites, this ratio is
always less than about 5, but materials with much lower aluminium
contents can be synthesized. The series of highly siliceous zeolites called
ZSM-5 with Si/Al typically 31 are a well-known range of catalysts with
extra stability conferred by the high silicon content. The 29Si spectra of
such substances are simple with essentially a single line due to Si(OSi)4
units. A second material isostructural with ZSM-5 but with only a trace
of aluminium and called silicalite is also known. If the aluminium
content is particularly low, then the single-line 29Si spectrum is found
to be resolved into a group of lines, which arise from the various crys-
tallographic sites in the as-yet uncertain structure of this material. This
remarkably well-resolved spectrum is also shown in Fig. 11.19. Because
the aluminium content of silicalite is so low, it was argued in order
to be able to patent its use, that it was not a zeolite and that any
aluminium was present as alumina impurity. 27A1 is a good, receptive
nucleus and can be detected at quite low levels in solids, which are
concentrated states of matter. The 27A1 MAS spectrum of the same
www.pdfgrip.com
Some applications 379
sample of silicalite is shown in Fig. 11.19, and it is evident that the 27A1
is detectable, even though an accumulation time of over two days was
required to collect the 176214 FIDs needed. The chemical shift is
diagnostic for tetrahedrally coordinated aluminium, so that the alu-
minium is to be found within the silicalite framework and, further, there
are at least two different aluminium environments. Note, again, that it
is the peak centroids that are indicated on the figure. The same type
of structure has since been observed in ZSM-5 that has been thoroughly
de-aluminated to reach Si/Al = 800.
An alternative approach to these catalysts is to take, for instance,
zeolite Y and subject it to what is known as decationation and ultra-
stabilization. The ammonium form of the zeolite is subjected to heat
treatment under vacuum conditions, when it loses ammonia and water.
The resulting crystalline material has much greater stability than the
starting material and is a good catalyst used for hydrocracking in
the petroleum industry. It has a much reduced ion-exchange capacity
and this indicates that aluminium has been lost from the framework,
the vacancies created being reoccupied by silicon. The aluminium
remains but can subsequently be leached out of the solid catalyst.
(a
(b) (C)
Figure 11.19 (a) The 29Si MAS spectrum of zeolite Na-Y with Si/Al = 2.61 obtained at 79.8 MHz. Five
Si environments are indicated, (b) The 29Si MAS spectrum of silicalite with Si/Al > 1000 obtained at
99.32 MHz. The resonances are all from SiO4 with no directly linked Al. (c) The 27A1 MAS spectrum of
the same sample taken at 104.2 MHz and the result of accumulating 176 214 FIDs. (From Klinowski and
Thomas (1985) Adv. Catal, 33, 199, and Fyfe et al (1982) /. Phys. Chem., 86, 1247, and Klinowski et al
(1984) Prog. NMR Spectrosc., 16, 237, copyright (1984) Elsevier Science, reprinted with permission.)
www.pdfgrip.com
380 High-resolution solid-state NMR
This ultrastabilization process has been studied by both 29Si and 27A1
spectroscopy, as shown in Fig. 11.20. The starting material (a) had
Si/Al = 2.61, four lines in the 29Si spectrum and all tetrahedral Al.
After calcining in air at 400°C for one hour (b), there are evidently
fewer aluminium atoms linked to the silicon and some octahedral
aluminium has appeared. Si/Al was calculated to be 3.37, a calcula-
tion that does not include the octahedrally coordinated metal, since it
is based on the 29Si intensities. More drastic treatment, heating in steam
at 700°C, produces even greater spectral changes and an Si/Al ratio
of 6.89 (c). Repetition of this procedure followed by leaching with
nitric acid (d) removes most of the aluminium to give Si/Al = 50 and
a single 29Si line, which indicates good crystallinity as is required if the
Al vacancies are filled. The octahedral aluminium resonance becomes
very narrow on leaching and represents remaining Al in the form of
[A1(H2O)6]3+ free to rotate in lattice cavities. These changes can also
be achieved by treatment with SiCl4 vapour or the aluminium can be
put back into the structure using A1C13 vapour, both processes having
been monitored by 27A1 MAS.
Si(2AI) Tetrahedral
\Si(1AI) (a)
Tetrahedral
(b)
Tetrahedral Octahedral
Si(OAI)
(c)
Octahedral
Si(OAI)l (d)
Tetrahedral
yy
_8Q -90 -100 -110 -120 200 100 0 -100
8 (ppm) from TMS 6(ppm)from[AI(H2O)6]3+
www.pdfgrip.com
Deuterium, an integral-spin nucleus 381
The patterns obtained in the solid state with the nucleus 2H, for which
7 = 1 , are somewhat different from those described above. A nucleus
with 7 = 1 has three energy levels and two degenerate transitions in
the absence of any quadrupole coupling. The interaction of the nucleus
with the magnetic field and the electric field gradient causes the three
energy levels to be modified, so that there are two transition frequen-
cies disposed symmetrically about the isotropic chemical shift value
and with a frequency difference that is proportional to the quadru-
pole coupling constant and to the orientation of the bond to deuterium
(and so the electric field gradient) relative to the magnetic field. In a
powder sample all orientations exist and the resulting 2H spectrum,
shown in Fig. 11.21, has a particular shape and is known as a Pake
powder pattern. The separation of the sharp edges of this spectrum is
three-quarters of the value of the quadrupole coupling constant and
usually lies in the range 120-150 kHz. If the moiety in which the
deuterium lies is capable of rotation in the solid, then the width of
the Pake pattern will be reduced and the extent of the reduction and
the shape of the pattern will depend upon the details of the motion.
The motion has to be fast relative to the value of the quadrupole
coupling constant. These comments will be illustrated by reference to
the 2H spectra of deuteriobenzene absorbed on graphite to form a
multilayer some ten molecules thick. Spectra were obtained at several
temperatures and are shown in Fig. 11.22. At 298K a singlet narrow
signal is observed, and shows that the absorbed benzene is reorienting
as if it were in the liquid phase. Indeed, this is found to be the case
even if the amount of benzene absorbed is reduced until it forms a
monolayer. At 170K the spectrum is of mixed form, with a Pake
pattern and a minor singlet. This latter is due to the absorbed mono-
layer, which is still undergoing fast two-dimensional motion, while the
Pake spectrum arises from benzene crystallites that form further out
from the surface. The splitting between the sharp edges of the Pake
pattern is 70 kHz, and this is typical of the value found for benzene
undergoing fast rotation around its hexad axis. The spectra obtained
at 90K were run using two different sets of conditions. In one, the
www.pdfgrip.com
382 High-resolution solid-state NMR
pulse repetition rate was slow, so that all the deuterons were detected.
In the other, the pulse repetition rate was much faster, so that the
component with the longer relaxation time T^ was saturated and effec-
tively was removed from the spectrum. In the first case, two compo-
nents are observed, both Pake patterns, and with splitting of 70 and
140 kHz. In the second case, the spectrum of the less mobile phase
has disappeared and the 70 kHz pattern remains. Note, though, that
this is not the same as that seen at 170K, since there is a distinct asym-
metry in the base. This is due to the now solid absorbed monolayer
rotating only around its hexad axis, and it follows that the broader
pattern arises from the now static benzene crystallites.
This type of spectroscopy is also much used in the study of liquid-
crystal phases, where the Pake-type patterns obtained from these
partially ordered materials can give much information about the
degree of order and the rates of motion, and how these change with
the experimental conditions.
50kHz
(a) (b)
50kHz
(c)
(d)
-v — «-v—
www.pdfgrip.com
Questions 383
11.7 QUESTIONS
11.1. Figure 11.23 shows the 13C NMR spectrum at 9.4 T of a solid
copper cyanide sample enriched in both 13C and 15N. The sample
is stationary. The shape of the spectrum shows principally the
chemical shift anisotropy though there are also contributions from
coupling (direct and indirect) between the nuclei present.
Calculate the approximate principal chemical shifts and the
isotropic chemical shift of the 13C given that TMS is at 0 ppm.
What information does this spectrum give about the structure of
(CuCN),7?
Figure 11.23 The 13C NMR spectrum at 9.4 T of solid copper cyanide enriched
in 13C and 15N. The sample was non-spinning. (After R.E. Waylishen et al.
(1999) /. Am. Chem. Soc., 121, 1528, with permission, copyright (1999)
American Chemical Society.)
50 0 -50 -100
ppm
www.pdfgrip.com
www.pdfgrip.com
Bibliography
www.pdfgrip.com
386 Bibliography
Braun, S., Kalinowski, H.-O., and Berger, S. (1996) 150 and More Basic
NMR Experiments: A Practical Course, 2nd edn, Wiley, New York,
ISBN 3527295127.
Canet, D. (1996) Nuclear Magnetic Resonance: Concepts and Methods,
Wiley, New York, ISBN 0471961450.
Derome, A.E. (1987) Modern NMR Techniques for Chemistry
Research, Pergamon, Oxford, ISBN 0080325149.
Fukushima, E. and Roeder, S.B.W. (1998) Experimental Pulse NMR,
10th edn, Addison Wesley, Reading, MA, ISBN 0201627264.
Martin, M.L., Delpeuch J.J. and Martin, GJ. (1980) Practical NMR
Spectroscopy, Heyden, London, ISBN 0855014628.
Mullen, K. and Pregosin, P.S. (1976) FT NMR Techniques-A Practical
Approach, Academic Press, New York, ISBN 0125104502.
Shaw, D. (1984) Fourier Transform NMR Spectroscopy, 2nd edn,
Elsevier, Amsterdam.
Neuhaus, D. and Williamson, M. (1989) The Nuclear Overhauser
Effect, VCH, Weinheim, ISBN 3527266399.
SHIMMING
DYNAMIC NMR
TEMPERATURE CALIBRATION
Van Geet, A.L. (1968) Anal. Chem., 40, 2227; ibid, 1970, 42, 679.
www.pdfgrip.com
Bibliography 387
RELAXATION
www.pdfgrip.com
388 Bibliography
Mann, B.E. and Taylor, B.F. (1981) 13C NMR Data for Organometallic
Compounds, Academic Press, London, ISBN 0124691501.
Pouchert, C.J. and Behnke, J. (1993) The Aldrich Library of 13C and
2
H FT-NMR Spectra, Aldrich Chemical, Milwaukee, WI, ISBN
0941633349.
Tebby, G. (1991) CRC Handbook of Phosphorus-31 Nuclear Magnetic
Resonance Data, CRC Press, ISBN 0849335310.
IMAGING
www.pdfgrip.com
Bibliography 389
The student may also care to read the following few original short
papers which summarize the early and unexpected results that
heralded the development of NMR as a subject useful to chemists:
Arnold, J.T., Dharmatti, S.S., and M.E. Packard, (1951) First obser-
vation of chemical shifts in a single chemical compound, /. Chem.
Phys., 19, 507.
Dickinson, W.C. (1950) Observed chemical shifts in fluorine
compounds and noted the effect of exchange, Phys. Rev., 77, 736.
Gutowsky, H.S. and D.W. McCall, (1951) An early observation of
spin-spin coupling, Phys. Rev., 82, 748.
Gutowsky, H.S., D.W. McCall and Slichter, C.P. (1951) A theory of
spin-spin coupling, Phys. Rev., 84, 589.
Proctor, W.G. and Yu, F.C. (1950) 14N chemical shift between [NH4]+
and [NO3],Phys. Rev., 77, 717.
www.pdfgrip.com
www.pdfgrip.com
Answers to Questions
Outline answers are given here. More detailed answers can be found
at http://www.thorneseducation.com
CHAPTER 1
CHAPTER 2
2.1 CHC12.
2.2 8 7.30, 2190 Hz, 8 7.30.
2.3 Worse.
2.4 -15950 ppm.
2.5 Ring current.
3537
2.6 C1 isotopomers.
2.7 The sign of the susceptibility correction is dependent on the
alignment of the magnetic field.
2.8 126.241 609MHz, 8981.
2.9 8 4519.7, 8 -204.0, 8 0, 8 -4701.5.
S-L _ Sll
2.10 ^^ ^ = Aref
y f - Asample*
y
www.pdfgrip.com
392 Answers to questions
CHAPTER 3
CHAPTER 4
4.1 22 s.
4.2 1 :1.15. Note !H and 13C relax 1H, whereas only 1H relaxes 13C.
4.3 Since the z coordinate is zero, then no terms can cancel and the
EFG is proportional to -3qr3. For zero EFG, each term of the
sum has to be zero, and if z is the coordinate of the displaced
charge, then 3z2 - r2 = 0 and z = ± r/V3.
4.4 46.9 s, 11.7 s.
4.5 1:0.103.
4.6 3.5 x 10-13 s.
14
4.7 N coupling. Cooling increases the rate of 14N relaxation and
decouples it.
4.8 Preferential rotation about the long axis. The sp carbon relax
by CSA, which depends on B02.
4.9 Scalar relaxation by 79Br.
103
4.10 Rh relaxes by CSA.
4.11 R2 = 282.74 s-1, T2 = 0.0035 s, R2SR = 280.51 s~\ T2SR = 0.0036 s.
www.pdfgrip.com
Answers to questions 393
CHAPTER 5
5.1 Dwell time = 56 jxs, 32K, 0.910 s, not fully relaxed, maximum
pulse width = 56 JJLS.
5.2 Mz, 0.368MZ.
5.3 Small error in frequency. Poor shimming.
5.4 4V2.
5.5 68.4°.
CHAPTER 6
CHAPTER 7
CHAPTER 8
www.pdfgrip.com
394 Answers to questions
CHAPTER 9
(Ph3P)2Pt
Mo(CO)3
(Ph3P)2Pt-p
Mo(CO)3
9.2 8(195Pt) = -4481; 8(3IP) = 22.1, 23.8; '/(195Pt,31P) = 3325, 3415 Hz;
8(195Pt) = -4449; 8(31P) = 22.9; 7(195Pt,31P) = 3325 Hz;
8(195Pt) (21.4) = -1607, -1575.
9.3 Five coordinate P group is chiral, hence C5H4R CH groups are
diastereotopic.
8(!H) = 4.73(2). 5.17(1), 5.23(1).
S(13C) = 92.3 (/ = 255 Hz); 85.8, 86.4, 94.6, 95.9 (/ = 15 for all).
9.4 P2 and P3 form an AB pattern centred on 30 700 Hz with 7AB =
490 Hz, /AM = 55 Hz, 7AX = 20 Hz, 7BM = 150 Hz.
3 195 1
9.5. /( Pt H) = 7.5,12 Hz. Viewed from platinum, the methyls are
inequivalent.
9.6.
2+
2+
www.pdfgrip.com
Answers to questions 395
9.9. H2 8 9.39; H5 8 8.38; H5' 8 8.30; H6" 8 10.1; H6'"8 6.5. Use COSY.
9.10. Three -1 : -1 :1 :1 quartets centred at 8 -414 (1), -420 (2), -422
(1).
9.11. yO^Rh'H) = 11 Hz produces 1 : -1 doublets. 7(lo3Rh31P) (PPh3)
= 109 Hz (triplet). '/(^Rh^P) {P(OMe)3} = 120 Hz (doublet).
CH,
(MeO)3P-Rh
L
'H
CHAPTER 11
www.pdfgrip.com
www.pdfgrip.com
Index
www.pdfgrip.com
398 Index
www.pdfgrip.com
Index 399
www.pdfgrip.com
400 Index
www.pdfgrip.com