+ All Categories
Home > Documents > NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes...

NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes...

Date post: 25-Dec-2019
Category:
Upload: others
View: 5 times
Download: 0 times
Share this document with a friend
119
NMR Spectroscopy in Structural Analysis Rainer Wechselberger 2009
Transcript
Page 1: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

NMR Spectroscopy in

Structural Analysis

Rainer Wechselberger

2009

Page 2: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

ii

This reader is based on older versions which were written and maintained by a number of people. To the best of my knowledge the following persons were involved in the history of this document: Rob Kaptein, Rolf Boelens, Geerten Vuister and Michael Czisch. The current version was completely revised by me and adopted to my lecture 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to improve the course material. Rainer Wechselberger, Utrecht in the summer of 2009

Please report errors in the text and/or explanations or any 'unclear' passages to:

[email protected]

or

[email protected]

Page 3: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

iii

I Introduction........................................................................................................ 1

1.1 Typical applications of modern NMR................................................................ 2

1.2 Some history of NMR ........................................................................................ 2

1.3 Aim of this course .............................................................................................. 3

1.4 General outline ................................................................................................... 4

II Basic NMR Theory ........................................................................................... 5

III An Ensemble of Nuclear Spins ...................................................................... 12

3.1 Ensemble of spins ............................................................................................. 12

3.2 Effect of the radio frequency (RF) field B1 ..................................................... 13

IV Spin relaxation ................................................................................................. 16

4.1 Molecular basis of spin relaxation................................................................... 17

V Fourier Transform NMR ............................................................................... 22

5.1 From time domain to spectrum ...................................................................... 22

5.2 Aspects of FT-NMR.............................................................................................. 26

VI Spectrometer Hardware.................................................................................. 28

6.1 The magnet........................................................................................................ 28

6.2 The lock ............................................................................................................. 30

6.3 The shim system................................................................................................ 31

6.4 The probe........................................................................................................... 32

6.5 The radio frequency system.............................................................................. 33

6.6 The receiver....................................................................................................... 33

VII NMR Parameters......................................................................................... 36

7.1 Chemical shifts.................................................................................................. 36

7.1.1 Effects influencing the chemical shift ................................................................. 38 7.1.2 Protein chemical shifts ....................................................................................... 39

7.2 J-Coupling......................................................................................................... 40

Page 4: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

iv

7.2.1 Equivalent protons ............................................................................................. 42

VIII Nuclear Overhauser Effect (NOE).............................................................. 44

8.1 Dipolar cross relaxation ................................................................................... 44

8.2 NOEs in biomolecules ...................................................................................... 47

IX Relaxation Measurements............................................................................... 50

9.1 T1 relaxation measurements............................................................................. 50

9.1.1 Calculated example for an inversion-recovery experiment................................ 52 9.1.2 Applications of T1 relaxation.............................................................................. 53

9.2 T2 relaxation measurements............................................................................. 53

9.2.1 Applications of T2 relaxation.............................................................................. 55

X Two-Dimensional NMR................................................................................... 56

10.1 The SCOTCH experiment ................................................................................ 56

10.2 2D NOE............................................................................................................. 59

10.3 2D COSY and 2D TOCSY ................................................................................ 62

XI The assignment problem ................................................................................. 66

11.1 Chemical shift ................................................................................................... 66

11.2 Scalar coupling ................................................................................................. 67

11.3 Signal intensities (integrals)............................................................................. 67

11.4 NOE data........................................................................................................... 67

XII Biomolecular NMR............................................................................................ 69

12.1 Peptides and proteins ........................................................................................ 69

12.1.1 Assignment of peptides and proteins ................................................................. 71 12.1.2 Secondary structural elements in peptides and proteins .................................... 75

12.2 Nucleotides and nucleic acid............................................................................ 79

12.2.1 Assignment of oligonucleotide and nucleic acid spectra................................... 85

XIII Structure determination .................................................................................. 86

13.1 Sources of structural information.................................................................... 86

Page 5: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

v

13.1.1 NOEs .................................................................................................................. 86 13.1.2 J-couplings ......................................................................................................... 87 13.1.3 Hydrogen bond constraints ................................................................................ 88

13.2 Structure calculations....................................................................................... 89

13.2.1 Distance-Geometry............................................................................................. 89 13.2.2 Restrained molecular dynamics ......................................................................... 91

Appendices.................................................................................................................. 94

Appendix A

a) Typical 1H and 13C chemical shift of common functional groups.............. 94

Appendix A

b) Random coil 1H chemical shifts for the 20 common amino acids............. 95

Appendix B: 1H chemical shift distribution of amino acids................................. 97

Appendix C: Nuclear Overhauser Effect .............................................................. 98

Appendix D: 2D NOESY experiment .................................................................. 101

Appendix E: Expected cross-peaks for COSY, TOCSY and NOESY for the

individual amino acids ................................................................... 103

Appendix F: Typical chemical shift values found in nucleic acid..................... 113

Appendix G: Typical short proton–proton distances for B-DNA....................... 114

Page 6: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

1

I Introduction

All spectroscopic techniques are based on the absorption of electromagnetic radiation by

molecules or atoms. This absorption is connected to transitions between states of

different energies. The nature of the electromagnetic radiation varies from hard γ-rays in

Mössbauer spectroscopy to very low energy radio frequency irradiation in NMR

spectroscopy. Other spectroscopic methods, applying electromagnetic radiation of

intermediate energy, are microwave spectroscopy (vibration and/or rotation of dipolar

groups in molecules), IR spectroscopy, where vibration states are excited, UV/vis

spectroscopy, where the electronic orbitals of atoms are involved and x-ray or atom

absorption spectroscopy involving the inner electron shells. In Nuclear Magnetic

Resonance (NMR) transitions occur between the states that nuclear spins adopt in a

magnetic field. Since the energy differences between these spin states are extremely

small, long-wavelength radio-frequency matches these differences. Accordingly, NMR is

a rather insensitive method and more sample-material is usually needed than for most

other spectroscopic methods. On the other hand, however, NMR lines are quite narrow

and therefore the resolution is usually so high that hundreds of lines can be resolved in a

single NMR spectrum. Also, the interaction between different nuclear spins is manifested

in NMR spectra, for instance, in the form of J-coupling or the nuclear Overhauser effect

(NOE). These properties have made NMR quickly an indispensable tool for structural

studies in chemistry and later also in biochemistry. NMR is the only available method to

date to determine the structure of proteins and nucleic acids in solution on an atomic

scale so that it became a well-established method in the field known as "Structural

Biology".

Page 7: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

2

1.1 Typical applications of modern NMR

Structure elucidation

Synthetic organic chemistry (often together with MS and IR)

Natural product chemistry (identification of unknown compounds)

Study of dynamic processes

Reaction/binding kinetics

Chemical/conformational exchange

Structural studies of biomacromolecules

Proteins, protein-ligand complexes,

DNA, RNA, protein/DNA complexes,

Oligosaccharides

Drug Design

Structure Activity Relationship (SAR)

Magnetic Resonance Imaging (MRI)

MRI today is a standard diagnostic tool in medicine.

1.2 Some history of NMR

NMR was discovered in 1945 by Bloch at Stanford and Purcell at Harvard University.

For this Bloch and Purcell received the Nobel Prize for physics in 1952. Initially, it

belonged to the realm of physics but after the discovery of the chemical shift (nuclei in

different chemical surroundings have different resonance frequencies) the technique

quickly became very important as an analytical tool in chemistry. The development of

Page 8: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

3

stronger magnets (maximum proton frequency is now (2006) about 1000 MHz) and of

multidimensional NMR methods allowed its entry in the field of biology. As a result of

its continuously increasing importance in modern chemistry, biochemistry and medicine,

two more Nobel prices for NMR followed in 1991 (Richard Ernst) and in 2002 (Kurt

Wüthrich).

1.3 Aim of this course

This course will bring the student up-to-date with the principles of modern NMR

methods and provide a basic understanding of how these methods work and how they can

be applied to derive the three-dimensional structures of biomolecules by NMR.

After a brief theoretical introduction of the basic physical principles of NMR, we will

discuss the origin of the parameters that determine the appearance of an NMR spectrum

such as chemical shift, J-coupling and line-width. Spin-relaxation (i.e. how a spin system

returns to equilibrium after excitation) is important as it determines the line widths of the

NMR signals, but also the intensity of the Nuclear Overhauser Effect, which in turn is the

major source of information for the structural analysis of biomolecules.

The modern way of recording NMR spectra is by applying short radio frequency pulses

and analyzing the response by Fourier transformation. This so-called Pulse Fourier

Transform NMR (for which R.R. Ernst received the Nobel prize for chemistry in 1991)

also allows the measurement of two-dimensional (2D) NMR spectra (and even 3D and

4D). An introduction is given to the FT-NMR technique and the principles of multi-

dimensional NMR are reviewed. Exemplary, some basic 2D NMR experiments will be

discussed in more detail. Finally, the important process of assignment of biomolecular

NMR spectra is explained and an overview of the possibilities to extract a (3-

dimensional) structure out of NMR data is given.

Page 9: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

4

1.4 General outline

I Introduction (what is NMR and what do you study with it)

II Theory (how does it work)

III Ensemble of spins (from single atom to real samples)

IV Relaxation I (after the experiment: back to equilibrium)

V FT NMR (with a single RF-pulse to a complete spectrum)

VI Hardware (what kind of device do you need for FT NMR)

VII NMR parameters (what you can see in an NMR spectrum and why)

VIII NOE (How does the Nuclear Overhauser Effect work)

IX Relaxation II (experiments to measure relaxation properties)

X 2D NMR (how to add an extra dimension and what's the good to it)

XI Assignment (which signal comes from which atom)

XII Biomolecular NMR (nucleic acids and proteins, spin systems and (structural)

parameters, sequential assignment)

XIII Structure Determination (which parameters to use, how to calculate a structure)

Page 10: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

5

II Basic NMR Theory

The energy states, between which transitions are observed during an NMR experiment,

are created only when a nucleus with magnetic properties is brought into an external

magnetic field. These magnetic properties of nuclei can be derived from a quantum

mechanical property, the spin angular momentum, I. Most nuclei have such a spin

angular momentum, which is represented by a corresponding spin quantum number I,

which can be integer or half-integer (I = 0, 1/2, 1, 3/2....) and we may simply speak of a

nucleus with spin I. The magnitude of the spin angular momentum is given by (2.1)

)1( += IIhI (2.1)

where h is h/2π ( h = 6.626 · 10-34 J·s is Planck's constant). Due to its quantum mechanic

nature, any component of I along an arbitrary axis of observation, for instance the z-axis

(which is by definition the direction of the external magnetic field), is quantized:

Iz mh=I (2.2)

mI is the spin quantum number which can adopt values from -I to I in steps of 1 (a total

of 2I +1 values). A combination of equations 2.1 and 2.2 leads to the spin-state diagrams

for nuclei with different spins (e.g. ½, 1, ³/2 ):

Page 11: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

6

A combination of their spin angular momentum and their positive charge causes nuclei to

have a magnetic moment (compare the effect of an electric current in a circular wire).

This magnetic moment is directly proportional to the angular momentum:

Iμ γ= (2.3)

γ is called the gyro magnetic ratio. Since I is quantized, accordingly also μ is quantized

and we can express μ in terms of the spin quantum number, I or μz in terms of the

magnetic quantum number, mI:

)1( += IIhγμ (2.4)

and

Iz mμ hγ= (2.5)

Classically, if we bring a bar magnet (compass needle!) in a magnetic field, denoted B,

the magnet will tend to turn and orient itself in the field. This is a consequence of the fact

that its energy, given by

Bμ ⋅−=E (2.6)

will then reach a minimum. For quantum mechanical objects such as nuclear spins the

situation is similar except that now only a limited set of discrete orientations (quantum

states) are available. Expression (2.6) still holds for nuclear spins. With the convention

that B lays along the z-axis we get the energy of a nuclear spin with a magnetic moment

of μ in an external magnetic field, B0 as

0BE zμ−= (2.7)

and with Eq. 2.5 it becomes clear, that the energy of a nuclear spin depends on the

magnetic quantum number mI:

0BmE Ihγ−= (2.8)

For a spin ½ nucleus this results in two states, denoted α for mI = +½ and β for

Page 12: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

7

mI = −½. As a consequence, the nuclear spins can not be found ‘turning around’ in order

to orient in an external magnetic field, but they rather can be found only in two different

orientations, corresponding to the two possible energy levels described by the two

magnetic quantum numbers:

In all forms of spectroscopy, transitions between energy levels are induced by

electromagnetic radiation of a particular frequency ν0, provided that the frequency

matches the energy difference between these energy levels:

0νhE =Δ (2.9)

This is sometimes called Einstein's equation and it is a basic relation in spectroscopy.

Based on Eq. 2.8 we find that ΔE = γhB0 = hν0, or

00 2B

πγν = (2.10a)

or even simpler:

00 Bγω = (2.10b)

Page 13: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

8

expressed in terms of the angular frequency

ω0 = 2πν0 (2.10c)

In NMR the relation (2.10b) is often called the "resonance condition" i.e. the condition

where the frequency of the radiation field matches the so-called Larmor frequency ωL =

γB0.

For nuclei with spin I larger than ½ we have multilevel energy diagrams. However, the

selection rule ΔmI = ± 1 still holds so that we arrive at the same resonance condition.

Since the energy of the spin states depends on the strength of the external magnetic field

(equation 2.6), we can modify the figure from above and adapt it for different magnetic

fields B0:

The separation ΔE of the energy states α and β and thus the resonance frequency,

depends on the sort of nucleus (γ) and the strength of the external magnetic field B0 or, in

other words, on the strength of the NMR magnet (compare 2.10).

E

B0

E = +½γ hB0 (β-state, m= −½)

E = −½γ hB0 (α-state, m=+½)

ΔE = γ h B0

B0

Page 14: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

9

Table 1 shows some properties of nuclei important for applications in organic chemistry

and biochemistry.

Table 1: Properties of selected nuclei

Isotope Nuclear spin

I

Resonance

frequency

(MHz)*

gyro magnetic

ratio γ [T-1 s-1]

Natural

abundance [%]

1H 1/2 600.0 2.6752 ∙ 108 99.985

2H 1 92.1 4.1065 ∙ 107 0.015

12C 0 - - 98.89

13C 1/2 150.9 6.7266 ∙ 107 1.11

14N 1 43.3 1.9325 ∙ 107 99.63

15N 1/2 60.8 -2.7108 ∙ 107 0.37

16O 0 - - 99.76

17O 5/2 81.4 -3.6267 ∙ 107 0.04

19F 1/2 564.5 2.5167 ∙ 108 100.0

31P 1/2 242.9 1.0829 ∙ 108 100.0

*resonance frequency at a magnetic field of 14.092 T (Tesla)

We will focus here on spins with I = ½ because they have only two possible energy states

and accordingly give only a single spectral line and thus are the most popular spins in

high-resolution NMR. For protein studies these are 1H, 13C and 15N (note that 12C has no

magnetic moment, and 14N has a spin 1). For nucleic acids in addition 31P is an important

nucleus.

Page 15: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

10

So far we have given a quantum mechanical treatment of the nuclear spin in a magnetic

field considering discrete energy levels. This led to the resonance condition (2.10).

Interestingly, a similar expression can be obtained from a classical description of the

effect of a magnetic field B on a spinning magnet with a magnetic moment μ that is tilted

with respect to the field direction. While a classical non-spinning bar magnet would just

orient itself in the direction of the field (compass), a spinning magnet cannot do this but

instead performs a precession about the direction of the field. In essence this is a

consequence of the conservation of angular momentum. There is a perfect analogy with

the motion of a spinning top (Dutch: "tol") in the gravity field of the earth. The spinning

top will also undergo a precessional motion. Mathematically, the effect of the torque

acting on a spinning magnetic moment is given by the cross product which can be written

as the determinant of a matrix. This is the “equation of motion”:

zyx

zyx

zyx

BBBdtd

μμμγγ

eeeμBμ −=×−= (2.11)

where ex, ey and ez are unit vectors forming an orthogonal

coordinate system along the x-, y- and z-axis, respectively.

Note that the direction of the precession is perpendicular to

both B and μ. With the usual convention that B is along the

z-axis (Bz = B0 and Bx = By = 0) the equations of motion for

the magnetic moment become

Page 16: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

11

0

0

0

=

−=

=

dtd

Bdtd

Bdtd

z

xy

yx

μ

μγμ

μγμ

(2.12)

It can be shown that a correct solution is given by

)0()(

)cos()0()sin()0()(

)sin()0()cos()0()(

00

00

zz

yxy

yxx

t

tBtBt

tBtBt

μμγμγμμ

γμγμμ

=

+−=

+=

(2.13)

This indeed describes a precession of μ about the z-axis with an angular frequency ω0

given by

00 Bγω = (2.14)

an expression identical with (2.10b). Thus, we have found that the quantum based

resonance frequency corresponds exactly with the equation for the classical precession of

a spinning magnet in a magnetic field.

In NMR we will often use a classical mechanics analogy for a description of nuclear

spins. Sometimes such an analogy must break down, however, since the spins really are

quantum mechanical in nature.

Page 17: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

12

III An Ensemble of Nuclear Spins

3.1 Ensemble of spins

In a sample of identical molecules we are dealing with a large number of nuclear spins.

In the quantum-mechanical picture these are distributed over the spin states according to

Boltzmann's law:

kTE

enn Δ−

β (3.1)

where nα and nβ are the populations of the

α and β state, respectively, and k is

Boltzmann's constant (k=1.3806504(24) ×

10−23 J/K). For I = ½ such a distribution is

shown on the right.

Due to the very small energy difference ΔE = γhB0 the

populations of the α and β states are almost equal. For a field B0

= 14 T (Tesla) (proton frequency 600 MHz) the relative excess

of α spins is only one in 104. This is one of the main reasons for

the low sensitivity of NMR spectroscopy.

In the (classical) vector model an ensemble of spins ½ can be

described as shown on the right.

The individual spin vectors all make an angle with the field B0

and are slightly more aligned parallel to the field than

antiparallel. In equilibrium the "phases" of the individual spins

Page 18: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

13

(their xy-components or positions) are randomly distributed. Therefore the resultant

magnetization vector M (the sum of all spin vectors) is aligned along B0. Like all the

individual spins that add up to M, also M has to be imagined

as spinning. So, if M is tipped away from B0 (see figure on the

left) it will perform a precessional motion about B0 with

frequency ω0 = γB0, just the way the individual spins do. This

‘tipping’ can be done by way of a radio frequency pulse - and

involves transitions between the spin states of the nucleus!

3.2 Effect of the radio frequency (RF) field B1

In all FT-NMR spectrometers an additional field B1 can be generated perpendicular to the

static field B0. This B1-field is created by means of radio frequency pulses (it’s basically

the magnetic component of the electromagnetic radio waves). What happens during the

pulse can best be compared to what happens to the spins, when they are brought into the

magnetic field of the spectrometer. We are just dealing with an additional magnetic field,

which tries to orient the spins in a new direction. The result of this additional field is a

rotation of the magnetization vector M around the axis of the additional field. This

rotation lasts only as long, of course, as the additional field is present. In other words:

only during the duration of the radio frequency pulse. In perfect analogy to the precession

around the B0-field, the speed of this rotation (its angular frequency) can be described as

11 Bγω = (3.2)

Acting on the equilibrium magnetization M, the effect of B1 is to tilt the vector away

from the z-axis. How far the magnetization is tilted depends on the sort of nucleus (γ), the

strength of the B1-field and the duration of the pulse. Illustrating the effect of a radio

frequency pulse is a bit tricky. Actually the magnetization vector is still precessing

Page 19: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

14

around the B0-field (with ω0) and at the same time, during the RF pulse, is tilted away

from the z-axis (is rotating around the x-axis with ω1)! Fortunately there is a simple way

to describe this: the rotating frame. In a frame of reference in which the x- and y-axes

rotate with ω0 about the z-axis, the B1-field appears stationary (left). In the rotating frame

the axes are denoted x', y' and z'. Now, at resonance, the motion of M becomes very

simple:

Rotating frame Laboratory frame

While in the laboratory frame the M vector performs a complex spiralling motion, in the

rotating frame it simply precesses about the direction of the stationary B1 (in the z'y'-

plane) with an angular frequency ω1 = γB1. Thus, by going over to the rotating frame of

reference we do not need to bother about the precession about B0 anymore. The effect of

an RF pulse can easily be described now: During the duration of a pulse (and only that

long!) the magnetization M is rotating around the axis from which the radio frequency

pulse is applied! The tilting of the magnetization vectors just follows the equation of

motion given in 2.13. If we keep the B1-field on for a short time such that

2 / t 1 πω = (3.3)

Page 20: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

15

and then switch it off we have tipped the magnetization along the y'-axis. This is called a

90° pulse. If we keep it on twice as long M is along the negative z-axis (180° pulse):

90° pulse 180° pulse

Remark: The concept of the rotating frame makes the description of NMR experiments

much easier. It is so convenient that we will use the rotating frame throughout the

complete course if not explicitly stated otherwise. For simplicity, we will use the notation

x, y, z instead of x', y', z' for the rotating frame in the following.

Page 21: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

16

IV Spin relaxation

After a 90° pulse from the x-axis M lies along the y-axis. If we leave the system now

undisturbed we will reach the equilibrium state again after a while by a process called

spin relaxation. Actually two things will happen:

i) magnetization will grow along the z-axis until the equilibrium value Meq has

been restored,

ii) the Mx and My components decrease to zero (usually faster than the equilibrium

magnetization is restored!).

The characteristic times for these processes are called T1 and T2. Thus we have

T1: longitudinal or spin-lattice relaxation time (z-magnetization),

T2: transverse or spin-spin relaxation time (x,y-magnetization).

In mathematical terms this can be described as follows:

(4.1a)

(4.1b)

(4.1c)

The solution of Eq. 4.1a is: [ ] 1)0()(Tt

eMMMtM eqzeqz

−−=− (4.2)

and of Eq. 4.1b: 2)0()(Tt

eMtM xx

−= (4.3)

2

2

1

)()(

)()(

))(()(

TtM

dttdM

TtM

dttdM

TMtM

dttdM

yy

xx

eqzz

−=

−=

−−=

Page 22: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

17

and similarly for My(t). Thus, all components of M return exponentially to their

equilibrium values. An important difference between T1 and T2 processes is that the

former involves changes in z-magnetization and hence, transitions between α and β spin

states that are accompanied by an exchange of energy with the "lattice" (environment). In

contrast, T2 processes involve loss of phase coherence in the xy-plane, no energy is

exchanged with the environment.

Note that spin relaxation is a random process and should not be confused with the

coherent rotations around B0- or B1-fields. For example, T1 relaxation which affects the

z-component does not create x- or y-magnetization (cf. Eq. 4.1).

4.1 Molecular basis of spin relaxation

What causes nuclear spins to relax? The simple answer is: exchange of energy with the

environment. But… the energies of the corresponding processes must match the ΔE

between the involved energy states. In most other spectroscopic techniques, this is

achieved by collisions (with other atoms or molecules). For the small energy differences

in NMR this is not an option. We have to look for processes with comparably small

energies (i.e. comparable frequencies) as the corresponding resonance frequencies in

NMR. We can find these in the interaction of moving magnetic dipoles. The magnetic

dipoles are the NMR nuclei themselves. The movement comes from the diffusional

motion of molecules. This process contributes both to T1 and T2 relaxation. But while

with T1 relaxation energy is exchanged with other molecules (the environment, the

‘lattice’) causing transitions between α and β states, in T2 relaxation the energy is

exchanged with spins of the same molecule, leading to a small variety in the precession

frequency of otherwise identical spins. The net-magnetization vector M is ‘split up’ in

many small components rotating with different frequencies. Eventually these components

are equally distributed in the xy-plane and no measurable transversal magnetization is

left. One speaks about dephasing of magnetization. It is important to note that also a

Page 23: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

18

static distribution of Bz-fields causes different frequencies (in different locations of the

sample) and also leads to dephasing i.e. T2 relaxation.

To understand the effect of motion, it is important to consider the time-scale of it. For T1

relaxation only motions with a frequency near the Larmor frequency ω0 are effective.

After all, they have to induce transitions between spin states and therefore must have a

frequency, which coincides with the ΔE between the states (thus, the Larmor frequency).

The time-dependence of the dipolar field comes from the rotational diffusion (the

‘thermal motion’, see figure to the left) of the molecules

which is characterized by a rotational correlation time τc.

For times smaller than τc the orientation of a molecule has

not changed much (leftmost figure below), while for t >>

τc the correlation between different orientations is lost

(rightmost figure):

For small fast tumbling molecules τc is quite short (10-11 - 10-10 s) while for large

biomolecules it is much longer (10-8 - 10-7 s). An approximate relation of τc with the

molecular volume V and the viscosity η is given by

TkV

cητ = (4.4)

For macromolecules of molecular mass Mr in H2O solution at room temperature a useful

approximation is

Page 24: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

19

12104.2

−≈ rc

Mτ (4.5)

In a randomly tumbling motion many frequencies are present. The distribution of the

frequencies of the motions is represented by the spectral density function, J(ω).

In this distribution the frequency ω = τc-1 acts much like a cut-off of the spectral density

function, in other words: motions with frequencies ω > τc-1 are quite rare. The area under

the curves is constant. Only the shape of J(ω) differs for molecules of different size. For a

smaller molecule for instance, J(ω) would look like:

τc for smaller molecules is shorter and their frequency distribution extends to higher

values of ω. On the other hand, J(ω) for low values of ω is relatively small. With

increasing size of the molecules J(ω) is getting larger for slow motions but at the same

time the ‘cut-off’ moves to lower frequencies. In other words, slow motions (small values

J(ω)

ω 1/τc

J(ω)

ω 1/τc

Page 25: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

20

of ω) are more likely to occur for (big) biomolecules than for small organic molecules.

Not very surprising indeed!

Now what does this mean for T1 relaxation? Here, the fluctuating fields have to induce

transitions between α and β states separated by the Larmor frequency ω0 = γB0.

Therefore, the efficiency of relaxation will depend on how much this frequency is present

in the distribution of frequencies of the molecule, thus on J(ω0). This is maximal for

molecules that have τc-1

= ω0, and the efficiency will drop for both larger and smaller

molecules (longer and shorter τc values). This explains the behavior of T1 versus

correlation time τc as shown in the next figure.

The minimum in T1 (most efficient relaxation) is at ω0τc = 1, which for common NMR

fields occurs for intermediate size molecules of molecular mass Mr ≈ 1000 D. In terms

of the fluctuating fields Bx(t) and By(t) a general expression for the efficiency (or rate) of

T1 relaxation is

Page 26: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

21

( ) )(10

222

1

ωγ JBBT yx += (4.6)

where the average of the square of Bx(t) , < Bx2 >, is a measure of the strength of the

fluctuating fields. Assuming that the components in all directions are the same ( < Bx2 >

= < By2 > = < Bz

2 > = < B2 > ) Eq. 4.6 becomes

)(210

22

1

ωγ JBT

= (4.7)

A similar expression for T2 relaxation is

( ))()0(10

22

2

ωγ JJBT

+= (4.8)

This describes the two mechanisms that contribute to T2: the static distribution of Bz-

fields (no or ‘zero’ frequency), J(0), and the effect induced by Bx(t) and By(t), which is

proportional to J(ω0). For larger biomolecules the J(0) term will dominate and becomes

approximately equal to τc. Thus, for slowly tumbling molecules we have the simple

expression

cBT

τγ 22

2

1 ≈ (4.9)

This means that T2 becomes progressively shorter for larger molecules and explains why

T2 unlike T1 does not go through a minimum.

Page 27: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

22

V Fourier Transform NMR

5.1 From time domain to spectrum

Nowadays, all modern NMR spectrometers work as so-called Fourier-Transform NMR

spectrometer (FT-NMR). Earlier we have seen (chapter 3.2) that the magnetic component

of an electromagnetic field (RF), applied along the x-axis of the rotating frame on

equilibrium z-magnetization, results in a precession around the x-axis when ωRF = ω0. For

the precession frequency, ω1, we found ω1 = γ B1. If we apply this RF field for a period t

= π/2ω1 we create 'pure' y-magnetization. An RF pulse of this duration was called a 90°

pulse. The RF transmitter of an NMR spectrometer is operated by a pulse-computer,

which can generate a single RF pulse or a series of RF pulses of arbitrary length,

frequency, phase, and amplitude separated by delays of adjustable length. An RF pulse of

length τp excites the frequency-range νRF - 1/(2τp) to νRF + 1/(2τp). To excite a certain

range of frequencies, τp must be adjusted to be sufficiently short. For example, at 600

MHz proton resonance frequency a good excitation of an NMR spectrum of 10 kHz

requires τp << 100 μs. In practice, pulses of τp = 2-20 μs are used.

A radiofrequency pulse of duration τp (left) and the corresponding excitation profile (right). Details in the

text above.

What will happen after a single RF pulse of 90° along the x-axis? Earlier we have seen

that the magnetization vector has been rotated and now is oriented along the y-axis (phase

τp

νrf

νrf-½τp

νrf

νrf+½τp

Δνrf

Page 28: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

23

coherence of the spins). The receiver is tuned to the frequency ωRF. If the resonance

frequency of spin j, denoted by ωj, is equal to ωRF the magnetization will remain along

the y-axis (of the frame rotating at speed ωRF). The magnitude of the magnetization will

be decreased by T2 relaxation. In case spin j has a resonance frequency ωj, which is

different from ωRF, then the magnetization

of spin j will precess in the frame rotating

at ωRF with the difference frequency

Ω = ωj- ωRF (5.1)

Needless to say, also in this case the

magnetization decays due to relaxation

processes. The in the xy- plane rotating

magnetization vector induces a current

when it passes the receiver coil. This is the

actual signal recorded by the receiver. In

order to be able to distinguish between

positive and negative frequencies (vectors

which are rotating clockwise or

counterclockwise with the same speed), both the x- and the y-component of the rotating

magnetization are recorded simultaneously. The corresponding signal induced in the

receiver coil has the shape of a decaying harmonics (sine and cosine waves) and thus is

called the free-induction decay (FID). For a number of different spins j (for example of

the Hα, the Hβ, and the HN), each with their own equilibrium magnetization Mjeq,

frequency ωj, and relaxation time T2j, the FID consists of the sum of all magnetizations:

[ ]

[ ] 2

2

sin)0()(

cos)0()(

Tt

Tt

etMtM

etMtM

jj

jx

jj

jy

′=

′=

ω

ω (5.2)

y

x

y

x

y

x

x component

y component

Page 29: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

24

where | Mj(0) | = | Mjeq | in the case of a 90° excitation pulse. The following figure

illustrates the difference in appearance of an FID containing only a single frequency and

an FID containing multiple frequencies:

The FT-NMR signal (the FID) is recorded in the time domain. The signal is digitized by

an analog-to-digital (ADC) converter and stored in the memory of a computer. The

resonance frequencies, ωj', are extracted by Fourier analysis. The Fourier Transformation

(Eq. 5.3) transforms the signals f(t) from the time domain to the frequency domain, g(ω):

dtetfg ti∫∞

∞−

−= ωω )()( (5.3)

where the complex signal f(t) is defined as

2)()()(Tt

ti eeMtMitMtf j

j

eqjxy

−′∑=+= ω . (5.4)

Fourier Transformation results in a sum of complex frequency signals. The real part of

Page 30: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

25

which describes an absorption signal:

22

2

2

)(1])(Re[

jj

j

j

eqj T

TMg

ωωω

′−+=∑ (5.5)

The imaginary part describes a dispersion signal:

22

2

22

)(1)(])(Im[jj

jj

j

eqj

TTMg

ωωωωω

′−+′−

=∑ (5.6)

The resonance line shape is a so-called “Lorentz line shape”. Usually we are only

interested in the real part of g(ω), the absorption response, because the flanks of

absorptive line drops with ω−2 , whereas the dispersive line goes with ω−1. This means

that the absorptive line is much narrower.

From Eqn. 5.5 we can calculate the

relationship between T2 and the line

width at half-height, Δν1/2:

2222 2

1115.0

νπ Δ+=

T (5.7)

Page 31: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

26

which can be rewritten as:

2

12

1Tπ

ν =Δ (5.8)

Some important Fourier pairs are shown here. Basically these are the ones responsible for

the shapes (or their distortion) of most lines observed on a FT-NMR spectrometer

5.2 Aspects of FT-NMR

It might be that the signal-to-noise ratio is not good enough after a single scan. By co-

adding n successive NMR measurements the signal, S, increases by a factor n. The noise,

FT

t

I

I

FT

FT

f(t) g(ω)

t 0

1

-1

t 0

1

-1

M

Page 32: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

27

N, however, fortunately increases with n only. This is due to the random nature of

noise. Hence the signal-to-noise ratio, S/N, improves as

nNS ~ (5.9)

FT-NMR is a very flexible technique. A large number of different experiments can be

done with the FT technique, each aiming at different parameters of the molecules in study

to be extracted, such as relaxation measurements (discussed in chapter 9), multi-

dimensional NMR (discussed in chapter 10), heteronuclear NMR, etc.

Page 33: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

28

VI Spectrometer Hardware

6.1 The magnet

The magnet is the core of the NMR

spectrometer. Nowadays mainly

persistent superconducting coils are used

to generate the high magnetic fields

necessary for high resolution NMR

(permanent magnets are only used e.g. in

food sciences or on older, lower field

NMR imaging systems). The coils consist

of NbTi (or NbTi-Nb3Sn, NbTi-

(NbTa)3Sn) wires which are

superconducting at 4 K (–269 °C).

Several layers of coils generate a higher

and higher field towards the innermost

section where the final magnetic field

strength is reached. In this section the

field must be extremely homogeneous over a volume of some cubic centimeters,

otherwise the resonance frequencies would vary at different locations in the sample

leading to broad and unsymmetric lines. The coil wires are also the most expensive part

of the spectrometer. To reach higher field strength, larger (and more complicated

designed) coils have to be used. This makes most of the price difference between e.g. a

700 MHz and a 900 MHz machine, which is approx. a factor of 4 (in 2006). The

necessary low temperature for superconductivity is reached by submerging the coils into

a dewar containing liquid helium (at –269 °C). This inner dewar is surrounded by a

second, outer dewar containing liquid nitrogen (–200 °C). In the course of time, both

Page 34: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

29

helium and nitrogen evaporate – therefore the ‘magnet’ (the dewars actually!) has to be

refilled periodically (typically weekly for N2 and monthly for He).

The strength of a magnetic field is normally given in Tesla or Gauss (1 G = 10–4 T). The

strength of an NMR magnet is often described by the corresponding resonance frequency

of hydrogen atoms (‘proton-frequency’). A field of 14 T corresponds to a field strength of

600 MHz. Today (2006) the typical field strength used in biological applications are 700

MHz and 800 MHz. The highest currently available field is about 1000 MHz.

The need for higher and higher fields is explained by the gain in resolution and in

sensitivity. The sensitivity of an NMR experiment is usually described by the signal-to-

noise ratio S/N:

S/N ~ N γ5/2 B03/2 n1/2 T2/T (6.1)

where N is the number of spins (concentration of the sample), γ the gyro magnetic ratio

of the nucleus, B0 the field strength, n the number of individual scans per experiment, T2

is the relaxation time and T the temperature of the detection circuit. How a bigger field

affects the S/N and the resolution (which follows a linear dependence on B0 ) is shown in

the following table.

B0 (T) 11.7 14.1 16.5 17.6 21.1

ν (MHz) 500 600 700 750 900

S/N 1.0 1.3 1.7 1.8 2.4

resolution 1.0 1.2 1.4 1.5 1.8

The Biomolecular NMR laboratory at Utrecht University is housing a 360 MHz, two 500

MHz, two 600 MHz, a 700 MHz, a 750 MHz and one 900 MHz high-resolution NMR

Page 35: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

30

spectrometer, one of the 600 MHz spectrometer and the 900 MHz spectrometer are

equipped with cryogenic probe systems for additional sensitivity.

6.2 The lock

Even in a very well designed magnet the magnetic field is not perfectly stable over the

long time a measurement can take (up to one week). Small deviations of the main

magnetic field can be compensated by the ‘lock system’ by applying correction currents

in a coil which is part of the room temperature shim system (see below). The lock system

exploits the NMR phenomenon itself: A reference NMR experiment is continuously

performed on a nucleus different from the one being studied. In most biological

Page 36: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

31

experiments deuterium (2H) is used for this purpose. The deuterium spectrum is

continuously acquired and the frequency of the single deuterium line is observed.

Whenever this frequency shifts, small correction currents are applied to the lock coil to

compensate for this change therefore slightly increasing or decreasing the total magnetic

field. In most biological applications deuterium is introduced by dissolving the sample in

a mixture of 5–10% D2O in H2O.

6.3 The shim system

As stated above, the magnetic field experienced by the sample must be very stable and

also very homogeneous to keep the lines as narrow as possible. Since the homogeneity is

not only a matter of the coil design, but is also influenced by the sample itself (filling

height, quality of tube etc.), for each individual sample additional field corrections have

to be applied. This is achieved by a number of correction coils (the shim system) in which

adjustable currents produce field gradients which can compensate field inhomogeneities.

There are two sorts of shim coils: superconducting ('cryo-shims') and room temperature

coils. The currents through the superconducting shim coils are usually only once adjusted

during the installation procedure of the magnet. The room temperature coils are the ones

used by the user for 'shimming' each individual sample. In practice the optimization of

the field homogeneity exploits the ‘lock’ experiment. The D2O in our sample gives a

single line in the NMR spectrum. The integral of this line is constant (as it only depends

on the number of nuclei in our sample, which is constant) but the height of the line is not:

The narrower the line the higher its maximum. The NMR operator can now manually

adjust the different currents in the different shim coils to optimize this value. This was

and still is a very time consuming procedure which requires some experience. Fortunately

a very fast automatic shimming method is available nowadays which employs pulsed

field gradients (PFGs) which reaches very good results within minutes. This so-called

'gradient shimming' is available on all of our machines (except the 360).

Page 37: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

32

6.4 The probe

The probe (or probehead) is in many ways the most critical component in the

spectrometer. It has two main functions:

a) to convert the radio frequency power from the amplifiers into oscillating magnetic

fields (B1-fields) and to apply these fields to the sample.

b) To convert the oscillating magnetic fields generated by the precessing nuclear

spins of the sample into a detectable electric signal that can be recorded in the

receiver.

Both points can be achieved by a parallel tuned circuit having a coil surrounding the

sample. The tuning is dependent from the sample (on position, volume, solvent, ionic

strength). The coil has to be carefully tuned to the frequency of the nucleus of interest

(remember: the frequency of the B1-field should match the Larmor frequency). This

adjustment of the circuit is important in several ways: First, we want to transmit the

maximum possible B1-field strength to the sample. This ensures that our pulses are as

short as possible, and therefore ensures a good excitation bandwidth. Second, since NMR

is a very weak phenomenon, we do not want to loose any signal coming from the sample

by picking up only a fraction of the oscillating magnetization.

There is a variety of NMR probes. For 1H spectroscopy typically probes are used which

can hold sample tubes of 5mm diameter (with a sample volume of ~500 μl). Beside the

proton channel there is another coil for the lock system tuned on 2H (sometimes a single

double tuned coil is used for both frequencies). The same setup can be found in probes

for 3mm and 10mm tubes. The sensitivity of the probes is still improvable as reflected by

the increased sensitivity over the past 10 years (nearly a factor of 2). This shows how

critical coil design is for NMR purposes. In a quite recent development, cryogenic probe

systems were introduced, which consist of a probe which can be cooled with cold helium

in order to reduce the amount of electronic noise in the receiver coils (and preamplifiers)

to a minimum. The technological challenge of such a system, among others, is the fact

that the temperature of the sample, of course, must still be adjustable to as high as about

80 ºC without heating the cold part of the probe. Since, on the other hand, the coils of the

Page 38: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

33

probe are supposed to be as near as possible to the sample, the difficulties of designing

such a system are obvious. The gains in sensitivity with the installation of such a system

to an existing spectrometer are remarkable and can be more than a factor of 2.2 (compare

the relative sensitivities at different fields in chapter 6.1).

6.5 The radio frequency system

The RF system mainly consists of pulse generation units, the actual transmitters and

subsequent power amplifiers. It is generating the excitation pulses at the frequency of the

nucleus of interest. On modern spectrometers the frequency can be set with a precision of

0.1 Hz across a band many megahertz in width. Since we may want to apply RF pulses

out of several directions in the rotating frame, the phases of the RF waves also must be

adjustable (typically to 0.5 degree). These settings are under extremely fast computer

control with setting times of only some microseconds. The power amplifiers boost the

transmitter output to high levels (from several tens up to hundreds of watts). This assures

that short, non selective pulses can be applied to the sample.

6.6 The receiver

The final stage in an NMR experiments is the detection of the precessing magnetization

(x- and y-components) in the sample. As stated earlier the same coil is used for this

purpose as for excitation. This means that directly before the data acquisition the

transmitter system has to be blanked and the receiver has to be opened (this ensures that

no strong RF pulses are applied while the sensitive receiver system is on). The detection

of the high frequency signal (MHz) is quite involved. First, analog filters are applied to

the signal to reduce it to the relevant frequency range. Then the weak signal is amplified.

The incoming signal is now mixed during several stages with reference frequencies. This

Page 39: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

34

mixing reduces the frequency of the signal from several MHz to the audio range (kHz).

Finally, the signal is digitized in real-time and stored. For digitization the following

relation is important: If we want to detect a certain spectral width SW we have to digitize

the signal with a time dw ('dwell time') or faster ('Nyquist theorem'):

SW

dw2

1= (6.2)

An example (see the following figure): Assume we digitize our FID every 5ms,

corresponding to a spectral width of 100 Hz. A frequency of 100 Hz will be sampled

twice per period (solid line), which is enough to characterize this frequency. On the other

hand, a resonance precessing with 120 Hz (dashed line) is sampled less than twice per

period, thus the frequency can not be distinguished from a slower frequency (80 Hz in

this case, dotted line). This means that both frequencies would give a signal at the same

position!

All further operations after the digitization and storing of the signal are performed in the

data processing system (the computer workstation). This includes application of window

functions, zero-filling, Fourier transformation, phase corrections, baseline corrections,

integrations in several dimensions as well as displaying and plotting the final spectrum.

Page 40: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

35

The components of a spectrometer at a glance:

RF generator (radio sender): creates the RF signal with a frequency of less than 100

MHz up to about 900 MHz.

Pulse generator: creates RF pulses of a duration of several µs up to several seconds.

RF amplifier: amplifies the pulse signal up to several 100 Watts.

Magnet: generates the B0-field (from about 1T up to about 21T).

Probe: holds the sample and houses the send and receive coils.

RF amplifier: amplifies the received signal from the probe.

Detector: Subtracts the base frequency from the signal, resulting in an audio frequency

(up to several kHz), containing only the differences of the resonance frequencies from the

base frequency.

AF amplifier: amplifies the audio signal.

ADC: Analog-to-digital converter.

Computer: controls all the other electronic parts, receives, stores and processes the NMR

signal.

Page 41: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

36

VII NMR Parameters

7.1 Chemical shifts

We have seen that the resonance frequency of a nucleus depends on its gyro magnetic

ratio γ and the magnetic field Bz. If all nuclei of the same kind (e.g. protons) would have

an identical Larmor frequency then NMR would not be a very useful technique for

studying biomolecules – we would observe just one line per sort of nucleus. Fortunately,

this is not the case since in practice different spins, even from the same sort, have a

slightly different Larmor frequency. This is because not all nuclear spins experience the

same effective static magnetic field Beff. Instead they experience the superposition of the

external field Bz and a local field Bloc. The static field Bz induces currents in the electron

clouds surrounding each nuclear spin. These induced

currents result in local magnetic fields. The induced

current will counteract its cause (‘Lenz law’,

electromagnetism), thus the induced field will be

opposed to Bz. The nuclear spins will be ‘shielded’

from the external field. The strength of this shielding

depends on the electron density around each

individual nucleus and the strength of the static field.

zloc BB σ−= (7.1)

where σ is a quantity expressing the amount of shielding. The net field experienced by

the spin becomes

)1( σ−=+= zloczeff BBBB (7.2)

Page 42: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

37

and the new Larmor frequency is given by (compare Eq. 2.10a)

π

σγν2

)1( zB−= (7.3)

The shielding σ is different for different types of nuclei in a molecule because the

electron density around a nucleus is very sensitive to the chemical environment of the

nucleus (e.g. chemical bonds and neighbours). The amount of shielding is usually given

as a dimensionless parameter δ, the chemical shift, which expresses the difference in

NMR resonance frequency with respect to a reference signal

)(101

10)(

10 666 σσσ

σσν

ννδ −⋅≈

−−

⋅=−

⋅= refref

ref

ref

ref (7.4)

since .1<<refσ

For the reference of the δ-scale the single line of the methyl-protons of Si(CH3)4 (TMS,

Tetramethylsilane) can be used (δ = 0). For biomolecules, slightly different compounds

(e.g. TSP, (CH3)3SiCD2CD2CO2Na, Sodium-salt of trimethylsilyl-propionic acid) are

used since TMS is not soluble in water; sometimes the water resonance itself is taken.

The chemical shift of water is temperature dependent:

9.96

][83.7)( 2

KelvinTOH −=δ (7.5)

The dimensionless δ-scale (ppm) has the important advantage over a frequency scale

(Hz) that the chemical shift values become independent of the magnetic field Bz.

Page 43: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

38

7.1.1 Effects influencing the chemical shift

As was mentioned above, the s-orbital electrons generate a field opposing the static field,

a shielding effect. The p-orbital and other orbital electrons with zero electron-density at

the nucleus result in a weak field reinforcing the static field. The contributions of these

two effects are of well known magnitude for the different functional groups. The table in

appendix Aa gives the common chemical shift values for a number of different functional

groups. A small value of δ corresponds to a shielded proton, or a low resonance

frequency.

In addition to the “constant” effects of s- and p-orbitals to the chemical shift, there are

also variable contributions resulting from the local surrounding of the nucleus (e.g.

solvent effects or interaction with other parts of the same molecule) and the local

conformation.

Aromatic and carbonyl groups have an extensive conjugated π-electron system

comprising delocalized molecular orbitals. Also in these systems the Bz-field induces

currents, the so-called ring-currents, resulting in quite large magnetic moments. Their

effect on δ of a particular nucleus depends strongly on the distance and orientation with

respect to the aromatic system: above and below the aromatic ring-system an opposing

field is generated and Δδ is negative (‘upfield shift’). In the plane of the ring-system a

reinforcing field is generated and Δδ is positive (downfield shift). These effects can be as

large as -2 to 2 ppm and all nearby protons are affected. The effect decreases with 1/r6.

Paramagnetic groups, like Fe3+ in the heme-system of hemoglobin, can have a

pronounced effect on chemical shifts. An unpaired electron influences the proton

chemical shift through spatial interactions, the electron magnetic moment, and by direct

electron-proton hyperfine interaction.

Electric fields resulting from charged groups or electric dipoles polarize the electron

clouds and thus influence the chemical shifts.

H-bond formation strongly influences the chemical shift. The proton in an X-H···Y

hydrogen bond is very little shielded and thus has a large δ value. For example, the imino

Page 44: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

39

protons in the Watson-Crick hydrogen-bonded bases of a B-DNA fragment resonate at

14-15 ppm, whereas the non hydrogen-bonded protons resonate at 10-11 ppm.

7.1.2 Protein chemical shifts

From section 7.1.1 it will be clear that also the 1H spectra of proteins will have a 'general'

appearance. An example 1H spectrum of a protein is given below:

Indicated are the typical regions where the different proton resonances are found. It is

even possible to split out the different chemical shift ranges on a per-residue basis. This is

shown in Appendices Ab and B.

Although the theory of chemical shifts is well known, it is quite complicated in practice

to accurately predict the chemical shifts in proteins. Partially this is the result of the

inaccuracy of protein structures and their internal mobility. On the other hand the range

Page 45: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

40

of proton chemical shifts is fairly limited (ca. 12 ppm) and the exact geometry is

relatively important.

The range of 13C chemical shifts is much larger (ca. 200 ppm) and the effects of the exact

geometry are less important. 13C chemical shifts are therefore easier to predict and can be

used in a more straightforward fashion for interpretation of spectra.

7.2 J-Coupling

J-coupling is the interaction between two spins transferred through the electrons of the

chemical bonds between them. The resonance frequency of a spin A depends on the spin-

state of a second spin B and vice versa. If spin A is in the α state spin B will resonate at

slightly lower frequency whereas it will resonate at slightly higher frequency when spin

A is in the β-state. In the NMR spectrum we observe a doublet (two lines with equal

intensity) centered on νb, the resonance frequency of spin B without J-interaction. Also

for spin A a doublet is observed, centered around νa. The size of the coupling is JAB (the

Page 46: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

41

distance between the components of the doublets) and is called the coupling constant. It

has the dimension Hz.

If there is a third J-coupled spin C the pattern splits again: the result is a doublet of

doublets. This means that the coupling of the third spin is independent of the coupling

between the first two spins. It just introduces another splitting on the first splitting. If the

Page 47: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

42

spins B and C are magnetically equivalent the doublet of doublet collapses into a triplet

(three-lines with intensity ratio 1:2:1). Three equivalent neighbours (e.g. the three protons

of a methyl group) result in a quartet (four lines with intensities 1:3:3:1). As you can see

the intensity ratios follow the famous Pascal triangle.

7.2.1 Equivalent protons

In general equivalent protons are protons which are chemically and magnetically

equivalent. Chemical equivalence means that there is a symmetry axis in the molecule for

the two protons under consideration. Nuclei having this type of equivalence resonate at

the same frequency. For magnetic equivalence the nuclei must be a) chemical equivalent

and b) must experience exactly the same J-coupling with all other nuclei in the molecule.

Magnetically equivalent nuclei are a very special case in the coupling network: They do

not couple with each other. This explains why for instance the benzene spectrum shows

only one line. Due to the high symmetry of the molecule all six protons are magnetically

equivalent, thus showing only one frequency and no coupling with each other.

In proteins magnetic equivalence due to symmetry is rare because of the high complexity

of the biomolecules. But also here some protons can be equivalent. If a group of atoms

rotates fast enough they become magnetically equivalent as a result of dynamic

averaging. This is the case for e.g. methyl groups or fast rotating aromatic rings.

Page 48: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

43

The magnitude of the J-coupling depends on the number of intervening chemical bonds,

the type of chemical bonds, the local geometry of the molecule, and on the γ values of the

nuclei involved. Proton-proton couplings in biomolecules are observed for protons

separated by two or three chemical bonds. The magnitude of these proton-proton J-

couplings is relatively small, typically 2-14 Hz. Often the patterns resulting from these 2JHH (two bonds) and 3JHH (three bonds) couplings are not resolved because of the large

line width in biomolecules.

The magnitude of heteronuclear one-bond couplings is much larger: the J-coupling

between the amide proton and the directly bond 15N nucleus, 1JNH is ca. 92 Hz. The one-

bond coupling between a proton and its directly attached 13C nucleus, 1JCH is ca. 140 Hz.

J-couplings involving 15N or 13C nuclei also exist between nuclei two or three bonds

apart. For example, there is an interaction between Hα and 15N over three bonds in a

HαC(C=O)15N fragment.

Page 49: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

44

VIII Nuclear Overhauser Effect (NOE)

The observable intensity of the signal of a nucleus depends on the intensities of other

nuclei when they are in close spatial proximity. If, for instance, two protons are situated

at a distance of less than 5 Å and the signal of one of them is saturated by selective

irradiation, the other signal will change in intensity. This effect is called the nuclear

Overhauser effect (NOE). It is the result of a relaxation process, caused by a dipole-

dipole interaction (dipolar coupling) between the two nuclei. We are talking here about

cross-relaxation, because the population of the spin-states of one nucleus depends on the

population of the spin-states of another one.

8.1 Dipolar cross relaxation

Let us consider a spin system with two spins

A and B which are dipolar coupled (spatial

proximity!). In the steady-state NOE

experiment one resonance is selectively

saturated by RF irradiation (let's say spin B).

This disturbs its equilibrium magnetization,

therefore spin B tries to re-establish it by

exchanging magnetization with its

environment. This can either be the lattice or

another nucleus close-by. The NOE between

the saturated spin B and another spin A is

defined by the relative change in the intensity

of spin A:

NOE = 1 + η with eqa

eqaa

MMM )( −=η (8.1)

Page 50: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

45

The energy level diagram for this two-spin

system is sketched on the right. Each of the

spins can undergo transitions between its α

and β state resulting in the resonance lines

which are observed. The rates of these

transitions are W1a and W1b for the

transitions of the spin A and B, resp. The

dipolar interaction between spin A and B

introduces two more possible transitions: W0 and W2. These transitions involve

simultaneous changes in the spin states of both the A and the B spin.

How this can be understood is schematically shown in the following figure:

The most left diagram represents the situation at equilibrium. A0 and B0 are the relative

population differences in equilibrium (corresponding to WA and WB). A and B are the

relative differences in population of the spin-states of the A and B nucleus after saturation

A

A

A

A

W2 > W0

small molecules

W0 > W2

large molecules

W1A

W1A

W1B

W1B

A0 = B0 = Δ

A = 1.5 Δ

A = 0.5 Δ

W2

W0

A

A

A = A0 = Δ

B = 0

Page 51: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

46

(diagram in the middle) and finally after cross-relaxation (diagrams on the right).

Saturation of the transitions of nucleus B leads to the situation in the middle. As a

consequence of the saturation, the population differences for the spins-states of nucleus B

disappear. Now the cross-relaxation comes into effect. Two different cases are

distinguished here: For small molecules, W2 dominates over W0 and the result is shown

in the upper right diagram. The relative population difference for the states of the A

nucleus has increased. Consequently also the observed signal for A will be increased. For

large molecules, W0 dominates over W2 and the result is depicted in the lower right

diagram. The relative population difference for the states of the A nucleus has decreased

and consequently also the observed signal for A will be decreased. One can easily

imagine a situation, where the two cross-relaxation mechanisms just cancel each other.

Indeed a zero-crossing of the NOE is observed when (in water at room-temperature):

ωτc ≈ 1.18

0.5

0.0

-0.5

-1.0

0.01 0.1 1.0 10 100

η

ωoτfast

tumbling

slow

tumbling

Page 52: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

47

Suppose for a particular molecule with a particular size on a particular NMR

spectrometer the 'observed' NOE turns out to be zero. Are there any options to still

observe NOEs within this molecule? Well, obviously we can do nothing about the size of

the molecule, but the tumbling speed of the molecule, of course, depends on the viscosity

of the solvent. The viscosity is usually very sensitive to changes in the temperature. So,

when we change the temperature of the sample, the tumbling speed will change and we

probably have a chance to pick up NOEs now. The other parameter which we probably

can change is ω0, which depends on the spectrometer frequency. So we could just repeat

the experiment on a spectrometer with a different field and chances are high that we

moved away from the zero-crossing situation. A more formal derivation of the NOE is

given in appendix C.

8.2 NOEs in biomolecules

T1 relaxation times are rather uniform for biomolecules. In contrast, cross-relaxation rates

vary a lot since the strength of the dipole field is strongly dependent on the distance

between the protons. A good approximation for the cross-relaxation rate, σ, in a

biomolecule is

6~rcτσ (8.2)

where τc is the rotational correlation time of the proton-proton vector, and r the distance

between the protons.

Since T1 values are uniform we can write for the NOE

6~rcτη (8.3)

Page 53: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

48

Therefore, the measured NOE can be converted to a distance. We can calibrate the NOE

by comparing it with a NOE of a fixed distance

refc

cref

ref rr

ττ

ηη ⋅= 6

6

(8.4)

If we now assume that there is no internal mobility (a rigid molecule), τc will be uniform

in the entire molecule. We can now directly calculate the distance

η refrefrr = (8.5)

Eq. 8.5 provides the basis of the most important effect for structure determination by

high-resolution NMR spectroscopy: the extraction of NOE-based distances between

proton pairs.

An example: In a protein we observe the following NOEs between the aromatic ring

protons and some other protons. The distance between CδH and CεH is fixed at 2.45 Å.

NOE intensity distance

Tyr CδH - Tyr CεH 0.15 2.45

Tyr CεH - Val CαH 0.02 3.43 r = 2.45 · (0.15 / 0.02)1/6

Tyr CεH - Asp CαH 0.01 3.85 r = 2.45 · (0.15 / 0.01)1/6

Internal motions are often fast (ωτc,intern < 1) and contribute predominantly to W2.

Page 54: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

49

Therefore, the NOE in mobile sub-domains will be reduced in intensity since

02 WW −=σ (8.6)

In the case of internal mobility we can not give the exact value of the distance but only an

upper limit. This is because of

6refc

crefrefrr

ττ

ηη

⋅⋅= (8.7)

and 1≤refc

c

ττ

(8.8)

Page 55: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

50

IX Relaxation Measurements

9.1 T1 relaxation measurements

The so-called inversion-recovery pulse

sequence, 180°-τ-90°-detection, can be

used for measuring the longitudinal

relaxation time T1. At the start of the

sequence, the equilibrium magnetization

Meq is inverted by a 180° pulse, after which the magnetization Mz(τ=0) = -Meq. The

magnetization will return to its equilibrium value ('recover') with the relaxation time T1,

and after time τ we have:

⎥⎥

⎢⎢

⎡−=

−121)(T

eMM eqz

τ

τ (9.1)

This time dependency is shown in the

figure on the right. We can measure the

value of Mz(τ) by applying a 90°

detection pulse, which will rotate the z-

component into the xy-plane. The FID is

recorded and the spectrum extracted by a

Fourier transformation. For τ < ln(2) T1

the spectrum is inverted. For longer values of τ, Mz(τ) has recovered to positive values

and a positive signal is recorded. By repeating the experiment with increasing values of τ,

the relaxation behavior can be determined and T1 extracted.

The T1 analysis is not limited to a molecule with a single resonance. In a molecule with

more spins, each of the individual spins, j, with frequency ωj and relaxation time T1j has

the starting magnetization:

Page 56: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

51

jeqzj MM −=)0( (9.2)

Therefore, we can determine the individual T1j by monitoring the intensities of the

individual resonances j at ωj in the spectrum as a function of τ.

The different time points τ of the inversion recovery sequence are measured one after

another. In case more than one FID is recorded per value of τ, e.g. for S/N improvement

or in multi-dimensional experiments (discussed in chapter 8), care has to be taken that

saturation is avoided and enough time is allowed between the individual experiments for

the magnetization to relax completely. The repetition rate of the experiment should be

adjusted so that at least a time of 5·T1 is waited before the experiment is repeated.

Page 57: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

52

9.1.1 Calculated example for an inversion-recovery experiment

We follow the intensity of one resonance and detect signal intensities between τ = 0.001s

and τ = 3s. Note that the inversion was incomplete as a result of an imperfect 180° pulse

( jeqzj MM −=)0( ). Therefore we have to use the more general Eq. 4.2 in place of Eq. 9.1:

[ ] 1

τ

)0()(T

eMMMM eqzeqz

−−=−τ

τ (s) Mz(τ) ΔM(τ)=Mz(τ)-Meq ln{ΔM(τ)/ΔM(0)}

“Mz (0)” 0.001 -12.0 -27.0 -

0.01 -9.4 -25.4 -0.06

0.03 -7.1 -22.1 -0.20

0.1 1.1 -13.9 -0.66

0.2 7.9 -7.1 -1.36

0.3 11.3 -3.7 -2.01

1.0 14.9 -0.1 -5.59

“Meq” 3.0 15.0 0

Since 1)0(

)(lnTM

M ττ −=⎭⎬⎫

⎩⎨⎧

ΔΔ

(9.3)

T1 can be determined by linear regression analysis:

Στ = 1.64; Σln(...) = -9.88;

and thus T1 = Στ / Σln(...) = 0.165 s.

Page 58: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

53

9.1.2 Applications of T1 relaxation

T1 tells us how long we have to wait until the equilibrium magnetization is restored. This

is important information for the setup of FT NMR experiments, where a number of

experiments are repeated and the results added up in the computer. It depends on T1 how

fast we can repeat our experiments.

9.2 T2 relaxation measurements

The value of T2 could in principle be extracted from the envelope of the FID or from the

line width at half-height (Eq. 5.8). However, T2 values obtained this way also depend

strongly on the static field inhomogeneities (the “shimming”). If the main magnetic field

is not homogeneous over the whole sample, spins at different locations in the sample tube

experience a slightly different field. This results in slightly different resonance

frequencies. The total peak, which is the sum over all individual spin contributions in the

sample, will be broadened due to this effect. The apparent relaxation time is T2* which is

faster than T2 due to spin-spin

relaxation. Since only T2 depends

on the physical properties of the

molecule, this is what we are

actually interested in. Obviously, the

contribution of a 'bad' shimming to

the relaxation of our molecule is less

interesting for us! 'Pure' T2 times can

be determined by the so- called

'spin-echo' pulse sequence, shown on the right.

The equilibrium magnetization, Meq, is transferred from +z- into y-magnetization, My, by

the 90°x-pulse. The macroscopic vector My can be considered to consist of a sum of

Page 59: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

54

macroscopic magnetizations Mj at j different positions in the sample. Now we look at the

rotating frame (which rotates with the average Larmor frequency ω0 of all the spins j): As

a result of an inhomogeneous field, we will find that some of the Mj components will

rotate faster than ω0 and some will rotate slower, depending on the exact position in the

sample. In the rotating frame (ω0), the fast and slow components will start to precess in

the xy-plane with frequency

jjz BMB Δ=− γωγ 0)( (9.4)

This frequency is different at different locations. Hence, the transversal magnetization My

will dephase due to the inhomogeneous field.

It can be shown that the spin-echo sequence eliminates the dephasing that results from

these static field inhomogeneities. In order to explain how this works, we actually should

consider the precession of each of the individual components Mj, but fortunately the

principle can be shown by picking only two spins, rotating with different speed: a slower

black and a faster white one.

Page 60: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

55

At a point (a) in the sequence we have pure y-magnetization for all individual spins. At

point (b) some phase coherence is lost because each spin has precessed with its own

frequency. The white one a bit faster, the black one a bit slower. After the delay τ/2 a

180° pulse from the y-direction is applied (point (c)). This pulse will invert the x-

component of mj, but will not affect the y-components. During the second delay τ/2, the

vectors (Mj) precesses again with their individual frequency γΔBj, the white one still a bit

faster and the black one still a bit slower and still in the same direction as before the

180° pulse. Consequently, in point (d) both magnetization vectors have returned to the y-

axis, independent of the static inhomogeneity, creating a so called 'spin-echo'. Thus, we

have eliminated the effect of field inhomogeneity! By repeating the experiment for

several values of τ in the range 0 - 4·T2 we can determine T2 from the decay of the

intensities of the resonances in the spectrum, which now results purely from the

relaxation by random fluctuating fields and is independent from any static field

inhomogeneity.

9.2.1 Applications of T2 relaxation

In chapter 4 we have seen that the T1 and T2 values depend on the motional behavior of

the dipole-dipole vector and thus the rotational correlation time τc of the molecule. Thus,

analogously to the usage of T1, we can use T2 to determine motional parameters.

Page 61: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

56

X Two-Dimensional NMR

In the seventies the development of two-dimensional (2D) NMR has revolutionized

NMR spectroscopy and has made the structural studies of biomolecules possible. The

basic idea is to spread the spectral information in a plane defined by two frequency axes

rather than linearly in a conventional one-dimensional spectrum. Clearly, this provides a

large increase in spectral resolution. Also, in a 2D NMR experiment interactions between

many spins in a molecule (whether it be J-coupling or NOE-type interactions) can be

measured simultaneously. This represents an enormous time-saving for large

biomolecules. To illustrate the method we will discuss first a very simple 2D NMR

experiment.

10.1 The SCOTCH experiment

SCOTCH stands for spin coherence transfer in (photo) chemical reactions. If one has, for

instance, a photochemical reaction

BA h⎯→⎯ ν (10.1)

a proton which resonates in molecule

A at ωA will resonate at ωB in

molecule B. The SCOTCH experiment

correlates the resonance frequencies in

A and B for each particular proton. In

other words it enables us to find ωB in

B that corresponds to ωA of the same

proton in A. The pulse sequence is as

Page 62: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

57

shown above. The 90° pulse creates xy- magnetization which precesses with ωA during

the so-called evolution period t1. The light pulse changes the precession frequency to ωB

with which it is detected as an FID during the detection period t2. The trick is now to

increment t1 in a regular fashion and collect a large number (typically, say, 500) FIDs

belonging to different t1 values. Thus one records a data set S(t1, t2) depending on both t1

and t2 (the FIDs). A first Fourier transformation 22 Ft → leads to a so-called

interferogram S (t1, F2) and after a second Fourier transformation 11 Ft → we arrive at

the 2D spectrum:

spectrum 2DraminterferogFIDs

),(),(),( 212121 FFSFtSttS FTFT ⎯⎯ →⎯⎯⎯ →⎯ (10.2)

The effect of incrementing t1 is to "sample" the frequencies present during the evolution

period. How this works out in practice is illustrated in Figure 10.1 (next page). After each

t1 time the signal acquired a different phase. The FIDs recorded that way have all the

same frequency (ωB), but their phase (i.e. how far the magnetization vector rotated in the

xy-plane) depends on the evolution time t1. After the first Fourier Transformation this

leads to a peak at the position ωB in the F2-dimension with an intensity that oscillates in

the t1 direction with the frequency ωA. Looking along the t1-axis of the interferogram the

signal actually looks like an FID oscillating with the frequency ωA. Hence, after double

Fourier Transformation this gives a peak at (ωA, ωB) (ωA in the F1- and ωB in the F2-

dimension). This is called a cross-peak, in contrast to peaks on the diagonal which are

called diagonal peaks. So, indeed, the experiment gives the connection (a spectroscopist

says: 'correlation') between the resonance frequencies in A and B. In this case this is

rather trivial but if there are many nuclei this method can prove very useful.

All 2D NMR experiments adhere to the following scheme:

Page 63: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

58

In the example above the preparation period would include a relaxation delay and the 90°

pulse. The mixing period is just the light pulse.

Fig. 10.1: The SCOTCH 2D NMR experiment. On the left side the sampling of the ωA frequency is shown at different values of t1. After Fourier Transformation of the FIDs (t2 → F2) the interferogram S(t1, F2) consists of lines at ωB in F2 with intensities dependent on t1 oscillating with ωA. The second Fourier transform (t1 → F1) leads to a 2D spectrum with a single cross-peak at (ωA, ωB). The representation is as a contour plot.

Page 64: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

59

10.2 2D NOE

2D NOE or NOESY is one of the

most important 2D NMR

experiments, because it measures

all short inter-proton distances in

a single experiment, for instance

for a protein. The first 90° pulse

belongs to the preparation period. The evolution time t1 is incremented and mixing

consists of two 90° pulses separated by a constant mixing time τm. During τm

magnetization between neighbouring spins is exchanged via cross-relaxation (see

Chapter 8). For biomolecules σAB ≈ W0 and therefore the flip-flop transitions (αβ →

βα) are dominant for the NOE effect. We will now look at the 2D NOE experiment in

more detail. Let us assume that there are two spins, A and B, within NOE distance, and

that the carrier frequency is chosen at the Larmor frequency of spin A, ωRF = ωA. The

vector diagrams at various times in the 2D NOE pulse sequence then look as shown in

Figure 10.2.

After the first 90° pulse the magnetization vectors lie along the y-axis in the rotating

frame. During the evolution time t1 the A-vector precessing at ωRF will stay the same (at

least for short times when relaxation can be neglected), while the B vector precesses with

a frequency ωB – ωRF. The second 90° pulse tips the A-vector to the negative z-axis and

the B-vector into the xz-plane. The z-component of B is shorter than that of A. We see

that the variable t1 time acts to create z-components with different magnitudes depending

on the different Larmor frequencies (i.e. how far a particular vector did rotate in the xy-

plane during t1. This is called frequency labeling and is a common feature of the

evolution period (t1 period) of 2D NMR experiments. Focussing now on these z-

components that correspond to populations of energy levels, we know that W0 transitions

during the mixing time τm will tend to equalize populations of αβ and βα states (because

Page 65: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

60

at equilibrium they are equal). Thus, the z-components of the A and B magnetizations

also will become more equal (d). Finally, the third 90° pulse flips the vectors in the xy-

plane where the signal can be observed (e).

Fig. 10.2. Vector diagram of the 2D NOESY experiment.

Now let us see how this leads to cross-peaks in the 2D NOE spectrum. In Fig. 10.3 we

shall look at the magnetization vectors at time d in Fig. 10.2 (after the mixing time τm).

Let us first consider a trivial case where no transfer occurs, for instance because the spins

are too far apart, and then the more interesting case where cross-relaxation occurs

between A and B.

Page 66: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

61

The vectors are depicted for various evolution times t1 chosen such that the B-vector has

rotated through 0, 90°, 180°, 270°, and 360°. If no mixing occurs the vectors precess at

their own frequencies ωA and ωB during t1 and continue to do this during the detection

period t2. Thus, this leads to a 2D spectrum after double Fourier transformation with only

diagonal peaks at (ωA, ωA) and (ωB, ωB).

Fig. 10.3. Vector representation of spin A (solid vector) and spin B (dotted vector) at time point (d) in Fig.

10.2 for various values of t1.When no mixing occurs during τm the 2D NOE spectrum consists only of

diagonal peaks. In the case of magnetization transfer during τm cross peaks arise at (ωA, ωB) and (ωB, ωA).

In contrast, when mixing occurs in τm the equalizing effect of the W0 transitions causes

the A-vector to borrow intensity from B and vice versa. Thus, the A-vector is now

modulated with ωB for the different values of t1! Since the vector will continue to precess

at ωA in t2 this will lead to peaks both at (ωA, ωA) and (ωB, ωA) in the 2D spectrum (F1,

F2). These are the diagonal peak that we have seen before at (ωA, ωA) and an off-diagonal

cross-peak at (ωB, ωA) at the upper left half of the spectrum. In the same way, the

Page 67: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

62

modulation of the B-vector with ωA leads to a symmetry related cross-peak at (ωA, ωB)

below the diagonal. Because mixing is reversible 2D NOE spectra are always

symmetrical. As all proton pairs within 5 Å will give rise to cross peaks with intensities

inversely proportional to r6, the 2D NOE spectrum provides a map of all short proton-

proton distances in a biomolecule. A more mathematical description of the NOESY

experiment is seen in Appendix D.

10.3 2D COSY and 2D TOCSY

In another important class of 2D

NMR experiments the magnetization

transfer in the mixing period takes

place via the J-coupling. The

simplest is the COSY (correlated

spectroscopy) with the pulse sequence shown

at the upper right. Here the mixing period is

just the second 90° pulse, which transfers

magnetization between A and B spins

whenever there is a J-coupling between them.

This results in the 2D spectrum as shown in

Fig. 10.4.

Fig. 10.4: COSY spectrum of two J-coupled nuclear

spins A and B. The sign of the cross-peaks is indicated.

The regular 1D spectrum is drawn above. It can be seen

that the cross-peaks reflect the fine structure of the 1D

spectrum (doublets in this case) and therefore can be

used to measure J-couplings. Note the typical negative-

positive sign pattern in the cross-peaks (but not in the

diagonal peaks).

Page 68: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

63

COSY spectra are often recorded at low resolution so that the fine structure is not visible.

In this way they are used to trace networks of J-coupled nuclei. Such low-resolution

COSY spectra are shown below for the two amino acids alanine and valine:

As measurable J-couplings only arise between nuclei separated by less than four

chemical bonds, the connectivity patterns can be easily predicted. We note that in a

protein a network of J-coupled protons does not extend beyond an amino acid residue

because the CαH of residue i is separated from the NH of residue i+1 by four chemical

bonds. Each of the amino acids forms a separate spin-system. A COSY spectrum is a

valuable tool for the identification of the types of amino acids.

Another important J-coupling based 2D experiment is the TOCSY which stands for total

correlation spectroscopy. The pulse sequence is shown below. The mixing period here

consists of a complicated pulse

train. This has the effect of

transferring magnetization

through a whole network of J-

coupled spins.

Page 69: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

64

For instance, in a molecular fragment

we have non-zero J-couplings 3JAB and 3JBC but 4JAC is close to zero. During the TOCSY

mixing period A-magnetization is transferred to B via JAB and then to C via JBC in

multiple transfer steps. Hence a cross-peak will arise between A and C even though there

is no direct J-coupling between these spins. To illustrate this, a comparison between

COSY and TOCSY spectra for the ABC fragment is shown in Fig. 10.5:

Fig. 10.5: COSY and TOCSY spectra of a three proton spin-system where JAB and JBC are non-zero and

JAC=0. Because of the symmetry only the part above the diagonal has been drawn.

It should be clear from this that a TOCSY spectrum always contains the COSY as a sub-

spectrum. Although the information content of COSY and TOCSY spectra is in principal

the same, in complicated spectra with a lot of overlap the TOCSY spectrum is still useful.

Page 70: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

65

This is because if B and C in the example of Fig. 10.5 are in crowded spectral regions but

A is not, then the whole spin system can be observed on a vertical line at the A-position,

while this would be difficult for B and C. We leave the sketch of the TOCSY spectrum of

alanine and valine as an exercise (compare Fig. 10.4 and 10.5).

Finally, it should be mentioned that the TOCSY is a sub-spectrum of the NOESY for

almost all cross-peaks. We will come back to this point in the following chapter.

Page 71: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

66

XI The assignment problem

The interpretation of an NMR spectrum always starts with the identification of resonance

frequencies and their corresponding nuclei in the molecule. This so-called ‘assignment’

of resonances constitutes an essential step in the structure determination process by high

resolution NMR spectroscopy that always precedes the actual calculation of structures.

While the assignment of smaller organic compounds with only a few 1H nuclei can often

be solved easily by means of a single experiment (e.g. COSY), it is much more

complicated for bigger and more complex molecules like peptides, proteins and nucleic

acids. Not only does the number of resonances increase with increasing size, also the line

width increases as a consequence of the shorter T2 relaxation times. As a result the lines

become broader and the overlap becomes increasingly severe.

We focus here on the basic principles of assignment, i.e. which parameters of the NMR

spectrum can be used. We will come back to the special case of the assignment of spectra

of biomacromolecules in the next chapter.

11.1 Chemical shift

The chemical shift of signals gives a first indication of the surrounding of the

corresponding nucleus in a molecule. We saw before (chap. 7.1.2) on the example of a

protein spectra, that proton chemical shifts usually are grouped according to the

environment in which they are located in the molecule. In our example those were the

amide and aromatic protons in the left half of the spectrum, the Hα protons just right of

the water signal (center of the spectrum), protons of aliphatic side chains still to the right

of them and finally the methyl groups on the right edge of the spectrum. If you look at the

vast variety of organic molecules you can find many more functional groups and

chemical environments a nucleus can be situated in. Most of the protons belonging to

such a group share their individual ranges of chemical shifts. Tables are available for 1H

Page 72: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

67

and 13C chemical shifts, which can help to identify the origin of particular resonances in

NMR spectra (see appendix A).

11.2 Scalar coupling

From chapter 7.2 we know that we can identify coupled nuclei with help of their coupling

constant. If we look, for instance, at the signals of the methyl group and the CH2 group of

ethanol, we find them split into multiplets due to scalar coupling. The number of

multiplet components gives us an indication of what the neighbouring group looks like

(i.e. the signal of the methyl group is a triplet due to the coupling with the two equivalent

protons of the neighbouring CH2 group. The signal of the CH2 group is a quartet due to

the coupling with the three equivalent protons of the neighboring methyl group). The

distance of the components of these multiplets (the coupling constant) is exactly the same

in the triplet and in the quartet. This helps to identify partners with a scalar coupling

between them.

11.3 Signal intensities (integrals)

The intensity, or better the integral of a signal tells us to how many equivalent nuclei a

particular signal corresponds to. The integral of the proton signal of a methyl group is just

three times as large as the integral of a single proton in the same molecule.

11.4 NOE data

NOE data can be very helpful for the assignment, especially when we already have an

idea of (basic) structural features of the molecule we are looking at. It is quite straight

Page 73: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

68

forward for example to identify neighbouring protons in an aromatic ring system, because

their distance from each other is quite short and well known (~2.45 Å). Once we know

one of them, we can relatively easily identify the others on hand of a NOE spectrum.

Especially in biomacromolecules, where the spin systems of the individual residues

cannot be connected by scalar coupling experiments, NOE data is often the only outcome

to still get a sequential assignment.

Page 74: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

69

XII Biomolecular NMR

Some of the most important sorts of biomolecules are nucleotides and amino acids.

Although also lipids and carbohydrates play important roles in biological processes, in

this course we will focus on the first two groups. Nucleic acid is the bearer of the genetic

information and is involved in protein synthesis where it acts as a template containing the

sequential information for all proteins occurring in organisms. Each three consecutive

nucleotides in a gene code for a particular amino acid in a protein. In addition there are

control regions (stop codons and sequences where proteins involved in transcription and

transcription regulation bind). The role of proteins is very diverse. We know them for

example as enzymes and regulators, as building material of cells and their compounds,

and as carrier of information within and between cells. Obviously both nucleic acids and

proteins play a major role in the function of all living organisms and accordingly also

most defective disorders of them can be traced back to the malfunction of proteins or to

defects in nucleic acid. This makes these molecules to very popular subjects of study in a

variety of research disciplines. Their structural and functional understanding is supposed

to give insight into how and why they work and what probably goes wrong in the case of

diseases. Knowing the exact composition and function of a particular virus, e.g. can lead

to the development of anti-viral drugs. The exact knowledge of the structure and function

of a particular enzyme can lead to the development of e.g. inhibitors which can deactivate

the enzyme when needed.

We will have a close look here at the structural properties of nucleotides and peptides and

how their different spin systems translate into different features in NMR spectra.

12.1 Peptides and proteins

Peptides and proteins are mainly build from twenty different naturally occurring amino

acids. They all share the same basic structure and only differ in their side chain R:

Page 75: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

70

H2N Cα CO

H

R

|

| N Cα CO

OH

H

R

|

|H|

H2N Cα CO

H

R

|

| N Cα CO

OH

H

R

|

|H|

In peptides, amino acids are linked via a so-called peptide bond. The amino group of one

residue is connected with the carboxyl group of another:

Note that the peptide bond is planar due to its partial double bond character! Amino acids

are usually referred to with either a one-letter or a three-letter code:

H2N Cα CO

OH

R

|

|Carboxyl group

Side chain

Amino group H2N Cα CO

OH

R

|

|H2N Cα C

O

OH

R

|

|Carboxyl group

Side chain

Amino group

MMetMethionineWTrpTryptophane

CCysCysteineYTyrTyrosine

RArgArginineFPhePhenylalanine

KLysLysineTThrThreonine

QGlnGlutamineSSerSerine

NAsnAsparagineIIleIsoleucine

EGluGlutamateLLeuLeucine

DAspAspartateVValValine

PProProlineAAlaAlanine

HHisHistidineGGlyGlycine

MMetMethionineWTrpTryptophane

CCysCysteineYTyrTyrosine

RArgArginineFPhePhenylalanine

KLysLysineTThrThreonine

QGlnGlutamineSSerSerine

NAsnAsparagineIIleIsoleucine

EGluGlutamateLLeuLeucine

DAspAspartateVValValine

PProProlineAAlaAlanine

HHisHistidineGGlyGlycine

Page 76: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

71

Amino acids can be classified by the character of their side chain as: aliphatic (A, V, L, I,

(G), aromatic/ring (F, Y, W, H, P), carboxylic (D, E, N, Q), sulfur/hydroxy containing (C,

M, S, T, Y) and charged (K, R). The chemical formulas of the natural occuring amino

acids together with their COSY, TOCSY and NOESY spectra are shown in appendix E.

12.1.1 Assignment of peptides and proteins

A strategy based upon homonuclear 2D experiments (COSY, TOCSY, and NOESY) was

developed in the 80’s (K. Wüthrich, 1986). This approach is discussed in this chapter.

A more recent approach employs uniformly 15N and 15N,13C labeled proteins. The

strategy uses so-called triple-resonance experiments (involving 1H, 15N and 13C) to

transfer magnetization through the polypeptide chain employing the large one-bond

homo- and heteronuclear J-couplings.

For larger proteins several patterns corresponding to residues of a certain type are present

in a COSY, e.g. several alanines give rise to similar patterns. How can we decide which

of the alanines in the protein sequence corresponds to a particular pattern?

The solution involves three steps:

1. Find patterns of coupled interconnected spins (spin-systems) belonging to amino

acid residues using COSY and TOCSY (amino acid identification).

2. Connect neighbouring spin-systems (in the sequence) with sequential NOEs.

3. Match stretches of connected spin-systems with the (known) amino acid sequence

for unique fits.

Let us focus now on the individual steps.

Page 77: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

72

Step1: Spin system identification

The identification of spins belonging to the same spin-

system can be performed on the basis of the COSY

and the TOCSY experiment (see the example of an

Ala-Ala peptide fragment on the right). In both

spectra, protons belonging to a certain amino acid can

be identified. A summary of all the expected patterns

for the different amino acid types is given in the

appendix E. Some of the amino acids have a very typical pattern, for instance Gly where

the side chain just consists of two Hα, or prolines where no HN is present. Some other

amino acids share a common pattern, e.g. the so-called AMX spin systems where AMX

represents the Hα and the two Hβ protons of the amino acid. To this group the following

residues belong: Phe, Tyr, Trp, His, Ser, Cys, Asp and Asn. In the COSY and TOCSY

spectra of Phe no J-couplings between Hβ and protons in the ring are observable (4 bonds

involved). This gap can be closed if in addition the NOESY spectrum is used since the

distance between these protons is typically smaller than 5 Å. Again, these cross peaks are

included in appendix E.

For bigger proteins the overlap makes it harder to identify such patterns. Also the lines

are broader in general which is due to the shorter T2 relaxation times. This also results in

a reduced efficiency of the magnetization transfer during the mixing period in TOCSY

since the magnetization is transversal during this time.

Step 2: Identification of neighbouring residues

When more than, for instance, one alanine is present in the protein, it is not a priori clear

which alanine in the primary sequence corresponds to a certain alanine pattern in the 2D

COSY spectrum. In order to make a so-called sequential assignment, i.e. correlating the

COSY patterns to individual amino acids in the primary sequence, we have to correlate

the COSY pattern of the alanine to the COSY pattern of its sequential neighbour.

Page 78: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

73

Unfortunately, when using only proton NMR, no 1H-1H

J-couplings of appreciable size exist over the peptide

bond since the shortest connection of two protons in

neighbouring residues involves four bonds.

Consequently, the individual amino acid residues form

isolated spin systems. Fortunately we can employ

another mechanism of magnetization transfer. The short

sequential distances between consecutive residues result

in cross peaks in the NOESY spectrum. The figure on the right summarizes the sequential

assignment approach: The type of the spin system is identified using COSY and/or

TOCSY experiments (bold arrows), whereas the sequential connectivity is established by

sequential NOESY cross peaks (dotted arrows). A cross peak between the Hα proton of a

spin i and the HN proton of the neighbouring spin i+1 results from a short distance

between these two residues, often referred to as dαN. Similar, the distances between Hβ or

HN of spin i to the neighbour HN of spin i+1 are represented by dβN and dNN .

On the other hand the tertiary structure of the protein will also lead to intense signals

from non-sequential cross peaks. How can we be sure to observe a sequential peak?

The statistics of these short distances have been investigated on the basis of thousands of

known (mostly X-ray) structures. Table 12.1 shows for example that 98% of all dαN

distances shorter than 2.4 Å correspond to sequential distances. Naturally, the score drops

with increasing distance limit. Similar values are obtained for the dNN and dβN distances.

This means that for two residues i and i+1 dαN, dNN and dβN cross peaks most likely result

from sequential NOEs. The probability for identifying a sequential connection increases

dramatically when simultaneously two (or more) short distances can be found.

Table 12.1 shows that if simultaneously two NOEs are found between two amino acids,

they most likely result from sequential residues.

Page 79: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

74

Table 12.1

Distance (Å) j − i = 1 (%)

dαN (i,j) ≤ 2.4

≤ 3.0

≤ 3.6

98

88

72

dNN (i,j) ≤ 2.4

≤ 3.0

≤ 3.6

94

88

76

dβN (i,j) ≤ 2.4

≤ 3.0

≤ 3.6

79

76

66

dαN (i,j) ≤ 3.6 && dNN (i,j) ≤ 3.0 99

dαN (i,j) ≤ 3.6 && dβN (i,j) ≤ 3.4 95

dNN (i,j) ≤ 3.0 && dβN (i,j) ≤ 3.0 90

Step 3: Matching to the sequence

The next step is to locate this fragment of two (or more) residues in the protein sequence.

For bigger proteins there might still exist several possibilities. Naturally, we can try to

link more patterns together and try to make tri-peptide, tetra-peptide, and even bigger

fragments. The uniqueness of such di-, tri-, and tetra-peptide fragments in proteins with

less than 200 residues has also been investigated. If all residue types could be identified

unambiguously in the fragment from the COSY (or TOCSY) spectra, a given di-peptide

fragment has a probability of uniqueness of 56%, the tri-peptide and tetra-peptide

fragments of 95% and 99%, respectively. As expected, increasing the length of the

fragment increases the uniqueness. A tetra-peptide fragment is usually sufficiently unique

Page 80: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

75

to allow for identification in the polypeptide chain.

12.1.2 Secondary structural elements in peptides and proteins

The intensities of sequential NOEs contain some information on the secondary structure

because they depend on the local conformation of the polypeptide backbone. Take, for

instance, the distance between an Hα proton of a residue i to the HN proton of the

following residue (i+1). In an extended strand this distance is short (2.2 Å) whereas it is

3.5 Å in a helical conformation. This information together with analysis of other

characteristic medium range NOEs (between residues which are less than 4 positions

apart in the sequence) is sufficient to specify the secondary structure element. An

overview of important sequential- and medium-range proton-proton distances is given in

the Figure 12.1.

Fig. 12.1: Characteristic sequential and medium range NOE connectivities.

Recognition of secondary structural elements, i.e. α-helices, β-sheets, and turns,

Page 81: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

76

constitutes an important element in the structure determination process. Most of the

observable NOE cross peaks between different residues are due to this secondary

structure.

The α-helix, β-sheet, and turn conformation result in characteristic short distances which

in turn result in characteristic NOEs in the 2D NOE spectrum. Also the 3JHNHα coupling

has traditionally been used as a marker for secondary structure as well as the presence of

slowly exchanging amide protons (cf. Chapter 13).

Fig. 12.2 shows the short distances in an anti-parallel β-sheet and in a parallel β-sheet.

Fig. 12.2: Anti-parallel (top) and parallel (bottom) β-sheet. Sequential NOEs are indicated by open arrows, interstrand NOEs by solid arrows. Hydrogen bonds connecting the strands are shown by wavy lines.

Page 82: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

77

The anti-parallel β-sheet is characterized by short dαN(i,i+1) and interstrand dαα(i,j)

distances whereas the dαα (i,j) distance is much longer (4.8 Å) in parallel β-sheet. Also

the dNN(i,j) distance in parallel β-sheet is much longer compared to the dNN(i,j) distance in

anti-parallel β-sheet.

In Fig. 12.3 short distances are shown for an α-helix. An α-helix is characterized by a

close proximity of residues i and i+3 and residues i and i+4. The dαN(i,i+3) and dαN(i,i+4)

NOEs are therefore clear markers for this element of secondary structure. In addition to

the aforementioned short distances, the sequential dNN(i,i+1) distance is also short, and

strong sequential dNN NOEs can be found in the spectrum.

Fig. 12.3: α-helix. The sequential dNN is shown (2.8 Å) together with dαN(i,i+1), dαN(i,i+2) and dαN(i,i+3),

Note that the side-chains are not shown.

Turns are characterised by short distances between residues i and i+2; in particular the

dNN(i,i+2) distance. In general, however, since there exists a large number of turns, which

Page 83: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

78

all have mirror images as well, e.g. I, I’, II, II’, III, III’, the precise nature of the turn is

hard to establish from NOE and J-coupling data alone. An overview of the characteristic

patterns and short distances is given in Fig. 12.4.

Fig. 12.4: Characteristic NOEs for several secondary structure elements. The thickness of the bars reflects the strength of the NOE. The thicker the bar the stronger the NOE (and the shorter the distance between the protons involved). 3J-coupling constants are also given.

The short distances in secondary structures are also listed in the following table:

Table 12.2: Secondary structure specific atomic distances (in Å)

Distance α-helix 310-helix β βp turn Ia turn IIa

dαN 3.5 3.4 2.2 2.2 3.4

3.2

2.2

3.2

dαN(i,i+2) 4.4 3.8 3.6 3.3

dαN(i,i+3) 3.4 3.3 3.1−4.2 3.8−4.7

dαN(i,i+4) 4.2

dNN 2.8 2.6 4.3 4.2 2.6

2.4

4.5

2.4

dNN(i,i+2) 4.2 4.1 3.8 4.3

dβN 2.5−4.1 2.9−4.4 3.2−4.5 3.7−4.7 2.9−4.4

3.6−4.6

3.6−4.6

3.6−4.6

dαβ(i,i+3) 2.5−4.4 3.1−5.1

a for turns, the two numbers apply for the distances between residues 2, 3 and 3, 4 respectively

Page 84: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

79

12.2 Nucleotides and nucleic acid

Nucleic acid is mainly build from five different nucleotides. All of them share a common

general structure: They consist of a nucleobase, a pentose-sugar ring (ribose in the case of

RNA or 2'-deoxy ribose in DNA) and a phosphate group which links the nucleotide units

to each other.

(Nucleo) Base

Phosphate

Sugar

Fig. 12.5: Common structure of nucleotides

There are two sorts of nucleobases: The purine bases adenine and guanine and the

pyrimidine bases uracil (only found in RNA), thymine (only found in DNA) and cytosine.

The base is connected to the sugar moiety via a glycosidic bond at the 1' carbon of the

pentose ring. In the common nucleotides the phosphate group is attached to the 5' carbon

of its sugar. A nucleotide which has no phosphate group is called a nucleoside. In

oligonucleotides and in nucleic acid, the phosphate group is linked between the 5' carbon

of one pentose and the 3' carbon of the next. It should be clear from this, that NMR

spectra of large nucleic acids are much more complex than NMR spectra of peptides of

comparable size. While peptides are build from as many as 20 different building blocks

(the different 'natural' amino acids) nucleic acid is build from only four different

nucleotides. Consequently the number of occurrences for a particular kind of nucleotide

is in average five times as high as for any particular amino acid. This can lead to very

crowded regions in the NMR spectra and can complicate the spectral assignment

considerably.

Page 85: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

80

1

2

34

5 6

7

8 9

1

2

34

5 6

7

8 9

1 2

3 4

5

6

Figures 12.6 to 12.9 show the different nucleobases and sugars with their numbering

schemes and the eight different RNA and DNA nucleotides.

Adenine

Uracil

Thymine

Guanine

Cytosine

Fig. 12.6: The five different nucleobases found in nucleic acids

Ribose (RNA) 2'-deoxy ribose (DNA)

Fig. 12.7: The pentose sugar rings of RNA and DNA

12

34

5

6

12

34

5

6

5'

4'

3' 2'

1'

5'

4'

3' 2'

1'

Page 86: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

81

AMP GMP Adenosine- Guanosine- 5'-phosphate 5'-phosphate

UMP CMP Uridine- Cytidine- 5'-phosphate 5'-phosphate

Fig. 12.8: Ribonucleotides (RNA)

dAMP dGMP Adenosine- Guanosine- 5'-phosphate 5'-phosphate

dTMP dCMP Thymidine- Cytidine- 5'-phosphate 5'-phosphate

Fig. 12.9: 2'-Deoxy-ribonucleotides (DNA)

Page 87: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

82

The figure below illustrates how the nucleotide units are linked by the phosphate groups

in oligo nucleotides and in nucleic acids. If we analyze the spin systems of this molecule,

we find that both the link between sugar and nucleobase and between the individual

nucleotide units (via the phosphate groups) reach further than three bonds before the next

proton can be found. In other words: The bases and the sugars form isolated spin systems.

This is important when it comes to think about an assignment strategy for these

molecules!

Fig. 12.10: The phosphate-sugar backbone of DNA

When we look at the structural properties of nucleic acids, the most prominent feature is

the double-helical conformation it adopts. The two strands of the helix adhere to each

other by means of hydrogen bonding between bases of the adjacent stands. These so-

Page 88: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

83

Geometry attribute A-form B-form Z-form

Helix sense right-handed right-handed left-handed

Repeating unit 1 bp 1 bp 2 bp

Rotation/bp 33.6° 35.9° 60°/2

Mean bp/turn 10.7 10.0 12

Inclination of bp to axis +19° -1.2° -9°

Rise/bp along axis 0.23 nm 0.332 nm 0.38 nm

Pitch/turn of helix 2.46 nm 3.32 nm 4.56 nm

Mean propeller twist +18° +16° 0°

Glycosyl angle anti anti C: anti G: syn

Sugar pucker C3'-endo C2'-endo C: C2'-endo G: C2'-exo

Diameter 26 nm 20 nm 18 nm

called base-pairs were found by James Watson and Francis Crick and accordingly named

'Watson-Crick base pairs'. Base pairs are always built from one purine and one

pyrimidine base. Adenine (A) always pairs with Uracil (U) in RNA and with Thymine

(T) in DNA. Guanine (G) always pairs with Cytosine (C). The hydrogen bonds are shown

in the figure below. Note that the G−C pair is more stable than the A−T pair, because

G−C consists of three hydrogen bonds and A−T only of two.

Fig. 12.11: Watson-Crick base pairs. A-T on the left, G-C on the right

There are two major conformation in which the double helices of DNA and RNA usually

are found. The so-called B-form and the A-form (see figure 12.12). Both of them are

right-handed (like ordinary corkscrews) and differ mainly in their width and height of

their turns. A third conformation, the Z-form is left-handed and of minor importance. The

important parameters of the A-, B- and Z-form helix are listed in table 12.3:

Page 89: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

84

A-form double helix (RNA) B-form double helix (DNA)

Fig. 12.12: The two major helical conformations

Information about typical chemical shift values of the A- and B-DNA and typical short

distances can be found in appendices F and G.

Page 90: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

85

12.2.1 Assignment of oligonucleotide and nucleic acid spectra

The assignment strategy for nucleic acid spectra is very similar to the one discussed for

peptides and proteins. The major difference being that for the identification of a particular

residue always the combination of COSY/TOCSY and NOESY is needed, because the

bases form isolated spin systems and cannot be assigned to their sugars without making

use of NOE data. Tables to be used for this purpose with common shift values and short

distances in nucleic acids can be found in appendix F.

Page 91: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

86

XIII Structure determination

In this chapter we will discuss the method of determination of the complete 3D structure

of a protein. Apart from the bond lengths and most bond angles, which are known on the

basis of the amino acid sequence, the most important source of structural information is

the NOE. In particular, the so-called "long-range" NOEs (those between protons more

than four residues apart in the sequence) provide important constraints on the structure.

We will first see which experimental NMR parameters can be translated into structural

constraints and then describe the computational structure calculation procedure.

13.1 Sources of structural information

13.1.1 NOEs

For a ~10kD protein typically between 1000 and 2000 cross peaks can be observed in a

2D NOE spectrum. We have seen in Chapter 7 how NOE intensities can be converted

into proton-proton distances with the aid of a reference distance (for instance the distance

of 2.45 Å between neighbouring protons on an aromatic ring). In principal the r-6 relation

between NOE and distance should give very precise distances. For example, if there is a

10% error on the NOE intensity this translates in only a 1.5% error in the distance!

However, there are two reasons why this high precision cannot be obtained in practice.

The first is local mobility. Remember that the simple relation of Eq. 8.5 was derived on

the assumption of equal τc for the protons of reference and unknown distance. Only for

very rigid proteins this assumption is really valid. Also, if there is a conformational

equilibrium the r-6 dependence gives values much more weighted to the short distances in

the average. Thus, the actual average distances may be longer than they appear in the

case of conformational averaging. The second reason is the so-called "spin-diffusion"

effect. This is the result of multiple transfer steps of magnetization A B C during

Page 92: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

87

the mixing time τm that may disturb the intensity of the NOE cross peak between A and

C. Only for very short τc can this indirect transfer path be neglected. For these reasons

the NOE based distance constraints are often used in the form of distance ranges rather

then precise distances:

strong NOE 1.8 − 2.7 Å

medium NOE 1.8 − 3.5 Å

weak NOE 1.8 − 5.0 Å

This procedure works well in practice because it turns out that for a high precision of the

structure a large number of NOE constraints is more important than precise distances.

13.1.2 J-couplings

The magnitude of the three-bond

coupling constant, 3J, is related

to the dihedral angle (between

the two outer bonds) and

therefore provides a constraint

on this angle (θ).

This is expressed in the Karplus relation

CBAJ ++= )cos()(cos2 θθ (13.1)

where A, B and C are parameters that depend on the particular situation. For instance, the

J-coupling between the amide proton and the α-proton, JHNHα depends on the backbone

angle φ as follows:

60.1)60cos(76.1)60(cos51.6 2 +−−−= φφJ (13.1a)

Page 93: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

88

The form of the Karplus relation for this case is shown in Fig. 13.1.

Fig. 13.1: The Karplus curve for the protein backbone φ angle as described in the text.

A complication is that several values of φ may belong to a particular J. Also,

conformational averaging may lead to average values of J in the range 6-7 Hz. Therefore,

the most reliable values for JHNHα are 9−10 Hz for extended (β-sheet) structure and 3−4

Hz for α-helical structure (Fig 12.4). Values around 6−7 Hz are often difficult to

interpret. Similar Karplus curves exist for CH-CH J-couplings from which side chain χ

angles can be derived.

13.1.3 Hydrogen bond constraints

When a protein is dissolved in D2O many of the amide protons do exchange rapidly with

deuterium and therefore disappear from the spectrum. However, often some NH signals

remain in the spectrum for some time and exchange only slowly in time. These slowly

exchanging NHs invariably are present in hydrogen-bonds such as occur in α-helices and

Page 94: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

89

β-sheets. If the H-bond acceptor is known with certainty, for instance, because we have

several short- and medium-range NOEs defining the secondary structure, we can use

distance constraints corresponding to the H-bonds. Usually these are given the distance

ranges 2.1 − 2.3 Å for the distance between NH and O and 3.1 − 4.3 Å between N and O.

13.2 Structure calculations

Several computer programs exist that are able to calculate the 3D structure of a protein

based on distance and dihedral angle constraints. Often one starts using only geometric

constraints (distances and angles) with the so-called Distance-Geometry (DG) program.

The resulting structures can then be refined with Molecular Dynamics (MD) calculations

which include also energy terms for electrostatic interactions etc. The calculation

procedure will now be briefly discussed.

13.2.1 Distance-Geometry

The structure of a macromolecule containing N atoms can be perfectly described by

specifying the N(N−1)/2 distances between the atoms (except for the chirality). Since this

is a very large number we will never have so many distance constraints in practice. There

is, however, an algorithm called Distance-Geometry (DG) that converts distances into

Cartesian coordinates for a much smaller number of distances even if they are not

precisely known. This algorithm has the following steps:

1. Set up distance matrices for upper bound (u) and lower bound (l) distances between all

atoms in the structure. These include the so-called holonomic distances (from bond

length and bond angles) and the NOE distance constraints. For those elements for which

no information is available the upper bound is set to a large value and the lower bound to

the sum of the van-der-Waals radii (1.8 Å).

Page 95: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

90

2. Smoothing

By this we mean the adjustment of upper

and lower bounds using triangular

inequalities. Consider three atoms i, j and k

with upper bounds uij, uik, and ujk, and lower

bounds lij, lik and ljk. The maximum distance

between i and j is when i,j and k are

colinear. Thus we have

jkikij uuu +≤ (13.2)

A similar relation exists for the lower bound lij:

jkikijl ul −≥ (13.3)

These relations are applied to all distances in the set until no further changes are obtained.

3. Select a matrix D with random trial distances between upper and lower bounds.

4. Embedding

This is the procedure of finding the best 3D structure that belongs to the distance matrix

D. We will not explain this further here.

5. Optimization

Because the matrix D, of course, is not perfect, it is usually necessary to regularize the

structure coming from the embedding stage. Optimization involves the minimization of

the coordinates against a distance error function.

Because of the random nature of step 3, the repetition of steps 3−5 will produce slightly

different structures. Ideally, if this is done many times, an ensemble of structures is

Page 96: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

91

obtained, which samples the conformational space consistent with the experimental data.

An example of such an ensemble of structures is shown in Fig. 13.2.

Fig. 13.2. Structures of the protein crambin calculated by the DG procedure.

13.2.2 Restrained molecular dynamics

Molecular Dynamics (MD) programs were originally designed to simulate the dynamic

behaviour of atomic or molecular systems. It turns out that a purely classical approach

works well for this purpose. For the initial coordinates and velocities of all atoms i,

Newtons equation of motion is solved:

ii

i dtdm Fr =2

2

(13.4)

where mi is the mass, ri the position vector and Fi the force acting on atom i. These forces

are given by

Page 97: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

92

i

iVr

F∂∂−= (13.5)

The empirical potential energy function V (also called the force-field) contains many

terms

⋅⋅⋅⋅+++++= ticelectrostaWaalsdervandihedralanglebondlengthbond VVVVVV (13.6)

that will not be specified here. For a faithful simulation of the dynamics the integration

step should be rather small, say 1fs (femto-second), so that for a dynamic trajectory of

several tens of ps or a ns very many integrations have to be carried out.

For the application of this method for NMR structure refinement we just add another term

in the potential energy function that reflects the NOE distance constraints:

ijijijij

ijijij

ijijijijNOE

lrrlk

urlururkV

<−=

<≤=>−=

if)(

if0if)(

2

2

(13.7)

This function looks as shown at the

right.

The effect of this term is the following:

If the distance rij in the current model is

too large (larger than the upper bound

uij) than a force is acting to decrease it.

Conversely, if rij is too small a force

will tend to increase it until rij lies

between the bounds. As this will happen for all 1000−2000 distance constraints the

structure will usually satisfy the constraints after a MD run better than before. At the

same time it will be close to a minimum with respect to the potential energy function of

Eq. 13.6. Thus, this so-called restrained MD algorithm acts as an efficient minimizer of

the energy. Of course, if the potential energy decreases the kinetic energy would increase.

Page 98: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

93

To avoid this 'heating effect' the system is normally coupled to a 'thermal bath' so that

excessive kinetic energy is drained off. An illustration of this procedure is shown in

Figure 13.3. Again, a family of structures is calculated starting with a random starting

structure. It can be seen that the final cluster of structures is better defined than those

from the DG calculations. This is usually expressed in terms of the root-mean-square-

deviation (RMSD) calculated for the ensemble of structures. Typically, RMSD values for

good NMR structures are in the range 0.3 Å to 0.5 Å for the backbone atoms.

Fig. 13.3: Calculation of the structure of dimeric interleukin-8 by a combination of Distance-Geometry and

Molecular Dynamics. The start is a random structure obtaining only the sequence information (only the

backbone is shown). 10 DG structures are calculated, which are the starting structures for the MD

calculation. After cooling down the precession of the structure is greatly increased.

Page 99: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

94

Appendices

Appendix A: a) Typical 1H and 13C chemical shift of common functional groups.

a)

Page 100: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

95

Appendix A: b) Random coil 1H chemical shifts for the 20 common amino acids

3.28 2.96 4.698.31 C

3.88 4.508.38 S 2.642.13

γ CH 2 ε CH 3

2.15 2.01 4.528.42 M

2.033.683.65

γ CH 2 δ CH 2

2.28 2.02 4.44P

7.156.86

7.307.397.34

1.23

1.480.950.89

1.640.94

0.97

H2,6 H3,5

3.13 2.92 4.608.18 Y

H2,6 H3,5 H4

3.22 2.99 4.668.23 F

CH 3 4.22 4.358.24 T

1.19CH 2 CH 3 CH 3

1.90 4.238.19 I

0.90H γ CH 3 1.65 4.388.42 L

0.94CH 3 2.13 4.188.44 V

1.39 4.358.25 A

3.978.39 G

H β H α NH

3.28 2.96 4.698.31 C

3.88 4.508.38 S 2.642.13

γ CH 2 ε CH 3

2.15 2.01 4.528.42 M

2.033.683.65

γ CH 2 δ CH 2

2.28 2.02 4.44P

7.156.86

7.307.397.34

1.23

1.480.950.89

1.640.94

0.97

H2,6 H3,5

3.13 2.92 4.608.18 Y

H2,6 H3,5 H4

3.22 2.99 4.668.23 F

CH 3 4.22 4.358.24 T

1.19CH 2 CH 3 CH 3

1.90 4.238.19 I

0.90H γ CH 3 1.65 4.388.42 L

0.94CH 3 2.13 4.188.44 V

1.39 4.358.25 A

3.978.39 G

H β H α NH

8.127.14

H2 H4

3.26 3.20 4.638.41H

1.703.327.17

1.451.703.027.52

2.386.87

7.596.91

2.312.28

7.247.657.177.247.5010.2

6.62

γ CH 2 δ CH 2 NH

1.89 1.79 4.388.27R

γ CH 2 δ CH 2 ε CH 2 NH 3 +

1.85 1.76 4.368.41K

7.59CH 2 NH 2

2.13 2.01 4.378.41Q

NH 2 2.83 2.75 4.758.75N

γ CH 2 2.09 1.97 4.298.37E

2.84 2.75 4.768.41D

H2 H4 H5 H6 H7 NH

3.32 2.99 4.708.09W

H β HαNH

8.127.14

H2 H4

3.26 3.20 4.638.41H

1.703.327.17

1.451.703.027.52

2.386.87

7.596.91

2.312.28

7.247.657.177.247.5010.2

6.62

γ CH 2 δ CH 2 NH

1.89 1.79 4.388.27R

γ CH 2 δ CH 2 ε CH 2 NH 3 +

1.85 1.76 4.368.41K

7.59CH 2 NH 2

2.13 2.01 4.378.41Q

NH 2 2.83 2.75 4.758.75N

γ CH 2 2.09 1.97 4.298.37E

2.84 2.75 4.768.41D

H2 H4 H5 H6 H7 NH

3.32 2.99 4.708.09W

H β HαNH

For X in GGXA, pH 7, 35ºC (Bundi and Wüthrich 1979)

Page 101: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

96

Random Coil Chemical shifts (in ppm) for the 20 common amino acids in acidic 8 M urea (from Wright,

Dyson et. al., Journal of Biomolecular NMR, 18: 43–48, 2000).

Page 102: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

97

Appendix B: 1H chemical shift distribution of amino acids.

Page 103: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

98

Appendix C: Nuclear Overhauser Effect

The population change of the αα-state, d/dt nαα, after a disturbance from equilibrium can

be expressed using the W0, W1a, W1b and W2 rates. The rate equation is:

)()(

)())((

11

2211

eqaab

eqaaa

eqeqba

nnWnnW

nnWnnWWWndtd

ββββ

ββββαααα αα

−+−+

−+−++−= (C.1)

Thus, the αα-state looses magnetization (first term), but also gains some magnetization

from the ββ-, the βα- and the αβ-states.

Similar expression can be found for the time dependence of the other three states.

Now we look at the net population differences of spin A and B, na and nb, resp. These are

)()()()(

βββααβαα

ββαββααα

nnnnnnnnnn

b

a

−+−=

−+−= (C.2a,b)

From this we can calculated the time derivate

βββααβαα

ββαββααα

ndtdn

dtdn

dtdn

dtdn

dtd

ndtdn

dtdn

dtdn

dtdn

dtd

b

a

−+−=

−+−= (C.3a,b)

If we now introduce in Eqs. C.3 the results from Eq. C.1 (and the expressions for the

other spin states) we can derive the dependency of the population from the transition

rates:

))(())(2(

))(())(2(

02210

02210

eqaa

eqbbbb

eqbb

eqaaaa

nnWWnnWWWndtd

nnWWnnWWWndtd

−−−−++−=

−−−−++−= (C.4a,b)

Page 104: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

99

The first terms of Eqs. C.4a and C.4b describe the T1 relaxation of spins A and B,

respectively. Hence we have

)2(1

)2(1

2101

2101

WWWT

WWWT

bbb

aaa

++==

++==

ρ

ρ (C.5a,b)

The second term of Eqs. C.4 results in transfer of magnetization from A to B. We define

the cross-relaxation σ between A and B as

02 WW −=σ (C.6)

With this definition, from Eqs. C.4 we arrive at the Solomon-Bloembergen equations:

)()(

)()(

eqaa

eqbbbb

eqbb

eqaaaa

nnnnndtd

nnnnndtd

−−−−=

−−−−=

σρ

σρ (C.7a,b)

Now let us return to the steady-state NOE experiment of chapter VIII. Spin B was

selectively saturated (nb = 0). After some time the two-spin system will reach a steady

state with

0=andtd

(C.8)

Eq. C.7a then becomes

)()(0 eqbb

eqaaaa nnnnn

dtd −−−−== σρ (C.9)

Page 105: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

100

so that the NOE (Eq. 8.1) becomes

a

b

a

b

a

a

aeq

eq

aeq

eqa

nn

nnn

γγ

ρσ

ρση ⋅=⋅=

−=

)( (C.10)

Here we exploited the fact that the macroscopic magnetization Ma is proportional to the

population na, and that the populations are themselves proportional to the gyromagnetic

ratio (Eqs. 2.8 and 4.1).

For identical nuclei γa = γb, and ρa = ρ, and Eq. C.10 reduces to the simple form

ρση = (C.11)

Page 106: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

101

Appendix D: 2D NOESY experiment

In mathematical terms the 2D NOE experiment can be described as follows. During the

evolution period the transversal magnetization of nucleus B can be written as

[ ])sin()cos( 111 titMeMM BB

eqeqB B

BB

tiωω

ω⋅+= = (D.1)

This results at time point c (Fig. 10.2) in a z-component

)cos( 1, tMM Beq

zB Bω−= (D.2)

According to the Solomon equation (Eq.C.7) we have for MA a dependency from MB

)()( eqBBAB

eqAAaA MMMMM

dtd −−−−= σρ (D.3)

For short mixing times 11

−=<< Am T ρτ we can neglect spin-lattice relaxation and Eq. D.3

becomes approximately

)( eqBBABA MMM

dtd −−≈ σ (D.4)

Using Eq. (D.2) this becomes

[ ])cos(1 1tMMdtd

BeqBABA ωσ +≈ (D.5)

For short mixing times τm we can approximate this by

[ ])cos(1 1tMMB

eqBAB

m

A ωστ

+=Δ (D.6)

Page 107: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

102

and thus a fraction of the A-magnetization ΔMA is modulated with ωB

[ ])cos(1 1tMM BeqBmABA ωτσ +=Δ (D.7)

During the detection period this evolves with 2ti Ae ω and after Fourier transformation we

will have a cross-peak at (ωB, ωA) in the 2D spectrum (F1, F2), correlating the protons A

and B in the 2D NOESY spectrum. The intensity of this cross-peak is

mABABI τσωω ~),( (D.8)

Since for biomolecules we have

6~AB

cAB r

τσ (D.9)

we get for the cross-peak intensity from Eq. (D.8):

6~),(AB

cmAB r

I ττωω (D.10)

Page 108: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

103

Appendix E: Expected cross-peaks for COSY, TOCSY and NOESY for the individual

amino acids. Diagonal peaks should be there, but are not shown!!

Page 109: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

104

Page 110: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

105

Page 111: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

106

Page 112: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

107

Page 113: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

108

Page 114: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

109

Page 115: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

110

Page 116: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

111

Page 117: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

112

Page 118: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

113

Appendix F: Typical chemical shift values found in nucleic acid

Page 119: NMR Spectroscopy · 'NMR Spectroscopy in Structural Analysis' I thank Hugo van Ingen, Marloes Schurink and Hans Wienk for proofreading, discussions, suggestions, and continuing to

114

Appendix G: Typical short proton–proton distances for B-DNA

All distances are given in Å. Sequential distances (to its 3' neigbor) below the diagonal,

intraresidual distances above the diagonal.

H6/8 3.8 2.1 3.6 4.1 4.9 3.4 4.4 4.8

3.5 1' 3.0 2.3 3.9 3.6 4.5 3.8

3.8 4.1 2' 1.8 2.4 3.9 3.8 4.0 2.1

2.3 2'' 2.7 4.0 4.9 3.6

4.9 3' 2.7 3.7 2.9 4.1

4.2 4' 2.7 2.3 4.9

1.8 4.3 3.2 4.5 3.6 5' 1.8 3.4

3.3 4.0 4.7 4.1 5'' 4.4

4.8 3.5 3.8 2.3 H6/8


Recommended