Вы находитесь на странице: 1из 24

Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

www.elsevier.com/locate/pnmrs

NMR methods for the determination of protein–ligand


dissociation constants
Lee Fielding *

Organon BioSciences, Newhouse, Scotland ML1 5SH, United Kingdom

Received 27 February 2007


Available online 24 April 2007

Keywords: NMR spectroscopy; Protein-ligand interactions; Binding affinity; Quantitative methods; Drug discovery

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
1.1. Ligand–protein interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
1.2. NMR of dynamic equilibrium states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
1.3. Accuracy and precision . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
2. Evolution of methods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
3. Protein observed chemical shift titrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
3.1. 1D 1H NMR studies of hevein domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
3.2. 1D 1H NMR studies of kringle fragments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
3.3. Other examples of 1H detected measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
3.4. Heteronuclear HSQC detected titrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
4. Ligand observed methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
4.1. Ligand observed chemical shifts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
4.2. Ligand observed relaxation rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
4.2.1. Ligand observed transverse relaxation rates 1/T2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
4.2.2. Ligand observed longitudinal relaxation rates (1/T1) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
4.3. Ligand observed translational diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
4.4. Ligand observed magnetization transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
4.4.1. Saturation transfer difference NMR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
4.4.2. WaterLOGSY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
4.4.3. Comment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
4.5. Ligand observed binding to serum albumins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
5. Competition binding experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
5.1. Graphical data treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
5.2. Magnetization transfer for sub micromolar binding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
19
6. F NMR studies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
6.1. Fluorine labelled proteins. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
6.2. Fluorine labelled ligands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
6.3. Fluorine observed competition binding experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
7. KD from ligand dissociation kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238

*
Tel.: +44 1698 736182; fax: +44 1698 736187.
E-mail address: l.fielding@organon.co.uk

0079-6565/$ - see front matter  2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.pnmrs.2007.04.001
220 L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

8. Alternative measures of protein–ligand binding affinity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238


8.1. Affinity index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
8.2. Affinity ranking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
8.3. NMR as a functional biological screen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
9. Applications of CP-MAS NMR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
10. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241

1. Introduction An earlier review [4] covered all of the early literature


pertaining to the determination of stability constants by
1.1. Ligand–protein interactions NMR methods. This earlier report describes the graphical
methods, which are now becoming only of historical inter-
Molecular recognition lies at the heart of biological pro- ests. Examples were drawn from the field of host–guest
cesses. At the molecular level the biological activity of drugs chemistry, but much is relevant to the study of the stability
corresponds to the binding of small molecules to macromo- of protein–ligand complexes. The monograph by Connors is
lecular receptors, usually proteins. This binding process is the definitive work on binding constants [5]. A binding
regarded as an equilibrium condition which results from a model is a prerequisite to any kind of data analysis. Usually
balance between association and dissociation events. The a 1:1 complex is assumed and this is very often taken for
quantitative foundations of pharmacology are mathematical granted. The classical NMR approach to the determination
models that describe these processes [1], and accordingly, of stoichiometry is the method of continuous variations
biological activity is expressed as the affinity of the partners (Job’s method) [6], but this is rarely applied to protein com-
for each other as a thermodynamic equilibrium quantity. plexes. In most reports the stoichiometry of the complex is
Binding affinities are usually determined in a binding assay, either well known from other studies, or it is safely assumed.
but increasingly in recent years, a large variety of physico- An advantage of using NMR to measure protein–ligand
chemical methods have been established for the quantitative interaction is that the NMR method extends the range of
determination of ligand–protein dissociation constants. measurable interactions into the mM range, a region not
This article looks at cases where NMR spectroscopy has well covered by traditional biochemical binding assays.
been used to determine the equilibrium binding constant
for small molecule–biomolecule complexes, with a focus 1.2. NMR of dynamic equilibrium states
on ligand–protein work. There are many diverse strategies
for making these measurements by NMR, so the main aim A protein and a ligand in thermodynamic equilibrium is
of the review is to discuss the available NMR experiments characterised by the dissociation constant KD, which for
and to illustrate, by means of examples the data analysis the simplest case of a protein with a single binding site is
methods that are applied. The variety of approaches to defined as
analysis of NMR data can obfuscate the underlying sim-
plicity of the methods. It should be noted that identical K D ¼ ½P½L=½PL; ð1Þ
experimental approaches and data analysis methods are where [P], [L] and [PL] are the equilibrium concentrations
used in other areas of chemistry, most notably in host– of protein, ligand and complexed state, respectively. [P]
guest chemistry. And, since such studies have been of inter- and [L] are also referred to as the free or non-bound states,
est since the earliest days of NMR in chemistry, much of e.g., ‘free ligand’, ‘non-bound protein’; and [PL] is vary-
what follows is not necessarily new. ingly referred to as ‘bound ligand’ or ‘saturated receptor’
An enormous amount has been published on NMR inves- depending on the context of the experiment. KD has the
tigations of protein–ligand interactions. As much again has units of concentration. Therefore a value of KD in the
been published on NMR studies of more general intermolec- mM range implies an approximately 1:1000 ratio of free
ular interactions involving species other than proteins. to bound states in an equimolar mixture of P and L and
Much of this literature deals with qualitative structural a KD in the lM range implies an approximately
issues. Useful earlier protein–ligand reviews that did address 1:10,00,000 ratio of these states, i.e., a much more stable
quantitative issues are the works of Dwek [2] and Gemmec- complex with less of the ‘free’ species present.
ker [3]. In order to have a clear focus and keep this review to In order to measure the dissociation constant of a pro-
a manageable size, the following discussion deals only with tein–ligand complex we have to measure the equilibrium
the quantitative analysis of protein–ligand interactions. concentrations of free and bound species. Traditionally this
The discussion does not aim to completely cover everything is done by equilibrium dialysis, but a great variety of other
that has been published on quantitative descriptions of pro- experimental approaches have been applied, including
tein–ligand interactions, but it does aim to be wide ranging NMR methods. To measure KD by means of NMR experi-
and it does comprehensively cover all possible approaches. ments implies quantitative analysis of solutions that are
L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242 221

potentially lM in the observed nucleus. The object of the 1.20


NMR observation might be the ligand, or the protein or 1.00

Mole fraction bound


both. In practice just one species is chosen as the observa-
0.80
tional target – the ligand or the protein and only very rarely
0.60
are both species viewed together. Hence, when viewing the
ligand, the NMR experiment has to be able to distinguish 0.40
between the free non-bound ligand and the ligand in the 0.20
bound state, so that [L] and [PL] can be quantified. Alterna- 0.00
tively, when observing the protein, the experiment has to dis-
-0.20
tinguish between and quantify [P] and [PL]. Protein–ligand
0 20 40 60 80 100 120
complexes are dynamic systems and therefore the rate at
[L] 0/[P]0
which the components of the protein–ligand complex
exchange between free and bound forms is central to the Fig. 1. Simulated speciation curves for protein (n) and for ligand (¤) for a
NMR method. series of solutions with [P]0 = 0.1 mM, and [L]0 is incremented through
0.01, 0.02, 0.05, 0.1, 0.2, 0.5, 1, 2, 5–10 mM. The assumed association
For a system in slow exchange on the chemical shift time
constant for a single binding site is 2 mM. The solution composition is
scale (in general terms this would be a protein–ligand system presented as the ligand:protein ratio. Mole fraction bound represents the
with high binding affinity and low dissociation constant, concentration of species in the complexed state relative to the total
KD = lM or lower), resolved signals might be expected for concentration of that species, XP(bound) and XL(bound) in the text.
the free and bound states, and using the ligand observed
example, [L] and [PL] are in principle available by integra- rated binding sites is what allows graphical solutions to the
tion of separate resolved signals. In practice this is extremely binding equations.
difficult because of the difficulty of detecting lM signals in Fig. 1 illustrates the formation of a 1:1 protein–ligand
what is likely to be a complex and crowded spectrum. complex for a hypothetical low affinity system with
For a system in fast exchange the observed NMR KD = 2 mM. The protein and ligand concentrations are
response of a ligand is the mole fraction weighted average presented as a ratio, as is commonly encountered. This is
of the NMR parameters of the free and bound states a typical example of NMR titration data. The XP(bound)
trace would result from observation of the protein NMR
M obs ¼ X LðfreeÞ M LðfreeÞ þ X LðboundÞ M LðboundÞ ; ð2Þ
signals, and the XL(bound) trace would result from observa-
tion of the ligand. The figure shows that the protein speci-
where Mobs is any NMR observable characteristic of the
ation varies considerably over the course of the titration,
equilibrium system, XL(free) and XL(bound) are the mole frac-
and suggests that a spectroscopic detector that was sensi-
tions of free and bound ligand, and ML(free) and ML(bound)
tive to this speciation would be a good means to measure
are the NMR parameters of the ligand in its free and bound
KD. The ligand speciation curve appears rather flat and
states, respectively. Similarly for observation of the protein,
unpromising as a measure of KD, but under appropriate
M obs ¼ X PðfreeÞ M PðfreeÞ þ X PðboundÞ M PðboundÞ ; ð3Þ conditions, ligand observations are just as effective.
The relationships between the relative responses of
where XP and MP are now the mole fraction and the NMR XP(bound) and XL(bound) become clearer when the solution
characteristic of the non-bound protein and occupied composition is plotted on a log scale and this is shown in
protein. The mole fractions are defined as, XL(free) + Fig. 2A. Moreover, the dependence of this speciation on
XL(bound) = 1 and XP(free) + XP(bound) = 1. KD is highly informative and this is shown in Fig. 2B and
Describing species distributions in terms of mole frac- C. It can be seen that as KD decreases (corresponding to
tions is a very useful concept and ‘mole fraction bound’ tighter binding) the protein and ligand speciation curves
is a term that occurs frequently in accounts of binding become more like each other and more like standard dose
interactions. It is generally used to describe the proportion response curves. XP(bound) always starts at zero and may
of a species present in the bound state relative to the total closely approach the limit of 1 providing that KD is small
amount. Both the ligand and the protein speciation can be or that the titration reaches high enough ligand concentra-
described in this way. So, whereas the mole fraction of tions to saturate the binding site. The starting value of
bound protein, XP(bound) might range from 0 to approach- XL(bound) is very dependent on the simulated conditions
ing 1 over the course of a titration with a ligand, the mole and in most practical cases will not even come close to
fraction of bound ligand XL(bound) might at the same time the limit of 1. XL(bound) will be approximately zero for
range from near 1 to approaching 0. Although these terms the duration of most NMR experiments. The usefulness
never reach their absolute limits (except for XP(bound) = 0 at of the mole fraction units is that this is what is measured
the start of a titration), in practice at experimental end by the equilibrium state analytical probe (the NMR data).
points they approach them closely enough that setting So, the process of measuring KD is one of transforming
X(bound) = 0 or 1 is a reasonable approximation. The mole fraction to concentration units (molarity).
almost complete approach of one of these parameters to Eqs. (2) and (3) are general for all NMR experiment
the end point conditions (0 or 1) under conditions of satu- observables. The NMR parameter may be the position of
222 L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

1.2 where [L]0 and [P]0 are the total concentrations of protein
1.0 and ligand, respectively. These total concentrations are usu-
Mole fraction bound

ally known. They represent the composition of the solution


0.8
as constructed by the investigator. The key to the determi-
0.6 nation of KD then, is to relate the known solution composi-
0.4 tion [L]0 and [P]0 to the equilibrium concentrations of ligand
0.2 and protein, [L] and [P], by means of the NMR observables.
Note though that the relationship between Mobs and KD
0.0
is non linear, and the parameter ML(bound) or MP(bound) is
-0.2
frequently not observable, so that it cannot be directly
-2 -1 0 1 2 3
log [L]0/[P]0 determined. There are two ways forward. Traditionally,
experiment protocols have been designed that linearize
1.2 the relationship between Mobs and KD. The data are then
1.0 analysed graphically and Mbound and KD come from the
Mole fraction bound

0.8
slopes and intercepts. Alternatively (and more often), the
relationship between Mobs and KD is solved computation-
0.6
ally and data analysis consists of determining a computed
0.4 best fit binding curve to the experimental data [7–9]. This
0.2 involves a two parameter fit to a binding curve. The bot-
0.0
tom line is that it is not possible to measure KD from a sin-
gle NMR data point and NMR protocols invariably
-0.2
-2 -1 0 1 2 3
involve monitoring some NMR signal as a function of
log [L]0/[P]0 varying solution composition. It is useful to talk not about
absolute peak positions or linewidths, but the changes to
1.2 these parameters that occur during a titration. Hence:
1.0
Dobs ¼ M obs  M free ; ð6Þ
Mole fraction bound

0.8
is the change in the NMR parameter induced by the equi-
0.6
librium condition relative to the free or non-bound state,
0.4 e.g., the induced chemical shift of a protein proton caused
0.2 by adding an interacting ligand, and:
0.0 Dmax ¼ M bound  M free ; ð7Þ
-0.2
-2 -1 0 1 2 3
is the limiting case of the above expression. Following the
log [L]0/[P]0 above example, for a protein reporter proton this would be
the chemical shift difference between the bound and non-
Fig. 2. Simulated speciation plots utilizing a logarithmic concentration bound forms. These two general terms, Dobs – the observed
scale and showing the effects of varying KD. (a) The same data as for
effect at any arbitrary equilibrium state, and Dmax – the lim-
Fig. 1, utilizing log [L]0/[P]0, KD = 2 mM. (b) and (c), Speciation plots
simulated for identical solution compositions with, (b) KD = 0.2 mM, and iting effect representing a fully bound state occur repeat-
(c) KD = 0.02 mM. edly throughout the following text.

1.3. Accuracy and precision


a resonance expressed in ppm or Hz, a linewidth, or either
of the relaxation rates 1/T1 or 1/T2, or any other feature of
The highest quality information on KD is in the curved
the system that reports on the dynamic averaging, e.g.,
regions of graphs such as Fig. 1 and in the steepest gradient
NMR measured translational diffusion coefficients. The sit-
region of Fig. 2. The regions of slowly changing mole frac-
uation of fast exchange also corresponds to the condition
tion (near to 0 and 1) are regions where the analytical
of the great majority of NMR based KD measurements,
response is insensitive to solution composition. Several
and so forms of Eqs. (2) and (3), written in terms of a spe-
detailed analyses of the errors associated with fitting spectro-
cific NMR parameter account for a large part of what
scopic data to binding models have appeared [10–14]. The
follows.
issues of concern revolve around obtaining spectroscopic
For the single site protein (1:1 complex), the solution
titration data that is actually responsive to the equilibrium
composition is defined as:
state and is not simply a response to binding; and how much
½P0 ¼ ½P þ ½LP; ð4Þ of a binding isotherm needs to be observed in order to sup-
and port the hypothesis of the binding model. The difference
between observing an equilibrium state and observing a
½L0 ¼ ½L þ ½LP; ð5Þ binding event is illustrated by many of the figures in the liter-
L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242 223

ature. Straight line regions of binding curves correspond to 2. Evolution of methods


either the progressive saturation of a species that is present
in sub stoichiometric amounts, or to a titration end point Approaches to the study of protein–ligand interactions,
where one species is fully saturated. Such regions contain and the measurement of binding affinities, have naturally
no information on KD. Information on the equilibrium con- been limited by whatever at the time was state-of-the-art
dition is only present in the curved middle region of the technology.
graph. The principal findings are as follows: The organization of this review broadly reflects the pro-
A ‘probability of binding’ (p) is defined as the ratio of con- gression of the implementation of methods. At the start
centration of complex to the maximum possible concentra- there was little choice. Only one or two methods were avail-
tion of complex. This formulation recognises that the able. In the 1970s, the only way that the stability constant
maximum possible concentration of complex is always the of a protein–ligand complex could be measured by NMR
initial concentration of the minor component (usually pro- was by means of 1/T2 or 1/T1 relaxation rate observations.
tein). A ‘saturation fraction’ has also been defined as the With the available 100 MHz spectrometers, there was little
ratio between the actual complex concentration and the ini- hope of resolving any protein signal. So the earliest NMR
tial concentration of the reagent, the chemical shift of which measurements of protein–ligand stability constants were
is being measured. This term is less useful for describing pro- based on relaxation rates of ligands in fast exchange with
tein binding situations because it does not reflect the fact the protein. Increases in spectrometer operating frequen-
that, at the start of the binding curve the concentration of cies coupled with the availability of pure protein and more
complex is limited by the concentration of added ligand. importantly, isotope enriched (15N and 13C) preparations,
The minimum error in the measurement of KD occurs at allowed direct protein observed titrations starting from
p = 0.5, and the ‘best’ data are obtained from the range the mid 1980s to the present. Now, twenty years later there
0.2 < p < 0.8. In other words, the most accurate values are many different tools available. Most recently, due to the
for KD are obtained when the equilibrium concentration development of a new range of ligand observed methods
of the complex is approximately the same as the free con- based on magnetization transfer effects, there has been a
centration of the most dilute component. return to ligand observed experimentation, with a lot of
The maximum information on the system comes from interest in saturation transfer difference NMR.
studying the widest possible range of p. At least 75% of Methods to deconvolute the concentrations of free and
the saturation curve is required in order to show correspon- bound species from exchange averaged NMR observations
dence between the equation of the model and the equation will be presented in Section 3 and early in Section 4. This
fitting the data (i.e., to verify that the binding model is represents the bulk of the published literature. The meth-
based on the correct stoichiometry). In other words any ods are conceptually easy to understand and easy to carry
binding data will be fit by any model over a suitably short out. Magnetization transfer difference experiments, which
range of p. If the experimental data are limited, higher effectively deconvolute the bound ligand fraction from
order complexes should be verified to be absent. Determi- the total ligand concentration within the exchange experi-
nation of the stoichiometry of a complex requires measure- ment, are the most up-to-date approaches and are consid-
ments at p = 1 (i.e., at undetectable protein or ligand ered at the end of Section 4. Most of the material in
concentrations). Since these conditions are the opposite Sections 5 and 6 can then be viewed as special cases and
of those required for an accurate measure of KD, the two as extensions of the general methods because no new
experiments should be separated. NMR detection is introduced. Some new ideas that work
The above comments cover the most important consider- best with 19F NMR are introduced at the end of Section
ations regarding experimental procedures. Some further rec- 6. The review ends with a consideration of several other
ommendations for the optimization of experimental diverse approaches that do not fit comfortably in the dis-
conditions for the determination of Ka, written in the context cussion so far.
of host–guest chemistry, but appropriate to this discussion NMR observations of protein–ligand systems are always
have been discussed [4]. These mostly relate to the graphical set up to observe either the protein or the ligand, never
methods. Wilcox has discussed these issues from the perspec- both together. So it is convenient to organize the following
tive of a more up-to-date NMR curve fitting context [15]. discussion according to whether the protein or the ligand is
This then is the reason that sensitivity considerations the observed NMR active species. One of the challenges in
limit the ability of NMR protocols to directly measure understanding the published literature in this area, is that
KD. Regardless of which NMR parameter is observed, of understanding the various data treatments that have
the experiment must be sensitive to analyte concentrations been applied. As discussed by Conners [5], there are three
that are approximately of the same magnitude as KD. In different ways in which to linearize the hyperbolic binding
other words, NMR techniques cannot directly measure a curve. To add further confusion, sometimes the protein,
value of KD smaller than the limit of detection of the exper- sometimes the ligand is at fixed concentration. Sometimes
iment. Typically this is around 10–100 lM for routine neither component is held constant and the data analysis
cases. The limit may be overcome by using competition applied may even be inappropriate. Add to this a few
experiments, as described in Section 5. unique and individual approaches to data analysis and this
224 L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

leads to many representations of graphs that illustrate the


derivation of KD from NMR titration data. different specialists does not take place. However
the aforementioned names have gained wide-
spread acceptance.
3. Protein observed chemical shift titrations
In the y-reciprocal plot (Scott plot) the ligand
concentration is plotted linearly along the
In this class of experiments, an NMR property of a
abscissa, thus retaining the scaling of the direct
nucleus in the protein is the marker of the binding equilib-
binding curve, but straightening the line. The
rium. Normally the response of this signal, perhaps the
NMR description of the bound state is obtained
chemical shift or linewidth of a protein proton, would be
from the reciprocal of the slope of the curve. The
monitored as ligand was titrated into the solution. It is
y intercept (crossing the ordinate at [L]0 = 0) gives
obvious that this method is only applicable to pure, solu-
KD/Dmax.
ble, modest molecular weight proteins and is subject to
In the x-reciprocal plot (Scatchard plot) the
all of the constraints that go with protein structure investi-
ratio of the NMR effect to ligand concentration is
gations by NMR. Isotope labelling is helpful, but not
plotted against the size of the NMR effect. The
essential. Titrating ligand into protein so that the ligand
slope of the line is the negative reciprocal of KD
eventually finishes in excess, thus saturating the protein
and the NMR property of the bound state is
binding site is the only way to perform this study. Little
obtained from the intercept with the abscissa.
useful information would come out of a protocol where
This data analysis has two distinct advantages
the ligand concentration never exceeded that of the protein.
over the others, KD is obtained from the slope
Neither will much useful information come from a system
of the curve, independent of any extrapolations,
where the ligand concentration vastly exceeds the protein
and the graph is ‘closed’ so that extrapolation
concentration unless the binding event is very weak (and
at each end leads to an intersection with an axis.
this case seems not to be interesting). Accordingly, the
The double reciprocal plot (Benesi–Hildebrand
experiments discussed in this section all follow protocols
plot) graphs the reciprocal of the NMR effect ver-
where the ligand:protein ratio is varied within a modest
sus the reciprocal of the ligand concentration.
range, 0.1–3 or 5. The linear graphical methods that occur
The slope gives KD/Dmax, and the intercept on the
widely in host–guest chemistry (see Box 1), are seldom if
y axis (1/[L]0 = 0), is the bound NMR parameter
ever applied to the study of protein–ligand interactions.
(1/Dmax). This data treatment has the advantage
Box 1 that it does not mix the dependent variable with
the independent variable (Dobs remains distinct
from [L]0).
There are three non-logarithmic linear solutions
Note that the protein concentration does not
to the NMR binding isotherm [5]. Written in terms
appear in any of these solutions. So there is no
of the NMR observables (for a titration of ligand
need for [P]0 to be accurately measured. The only
into protein, and where the protein is the focus
requirement is that [L]0 > [P]0. All three data treat-
of the NMR observation), they are
ments are likely to be found in the literature, but
½L0 =Dobs ¼ ½L0 =Dmax þ K D =Dmax ; ð8Þ with the Scatchard and Benesi–Hildebrand
Dobs =½L0 ¼ Dobs =K D þ Dmax =K D ; ð9Þ names occurring more frequently than Scott.
The Scatchard plot is generally preferred
1=Dobs ¼ K D =Dmax ½L0 þ 1=Dmax : ð10Þ
because KD comes without extrapolation to a
The graphs of these solutions are often referred potentially remote axis intersection.
to by the names of early researchers, so that the
graph of Eq. (8) is a Scott plot, the graph of Eq. (9)
is a Scatchard plot, and that of Eq. (10) is the Bene- The examples that follow are all based on the 1:1 bind-
si–Hildebrand plot. Connors terms them as a y- ing model, and an assumption that the system is in fast
reciprocal plot, an x-reciprocal plot and a double exchange. The case of insufficiently rapid exchange
reciprocal plot, respectively, according to whether between the two species being observed in the titration
the dependent parameter (y the NMR observa- (protein and bound protein) has been examined by Sudme-
tion), the independent parameter (x the ligand ier et al. [16].
concentration), or both appear in reciprocal terms. An advantage of protein observed NMR titrations, and
This terminology has the advantage that it identi- something that sets this method apart from ligand observed
fies the plot more precisely, and avoids the confu- NMR titrations, is that it is often possible to directly observe
sion caused by application of yet more names, as the fully bound state. That is, a protein spectrum may be
occurs when parallel developments are made in obtained under conditions of a fully saturated binding site.
different fields, but communication between When this is the case, Dmax (corresponding to dbound) can
be measured directly, and at the risk of introducing some
L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242 225

greater variance in the result, one unknown drops out of the fully saturated state (dbound), and at the equilibrium condi-
binding isotherm. Most workers prefer to treat dbound as an tion (dobs) according to:
unknown and find the best fit of both parameters to a binding
X PðboundÞ ¼ ½PL=½P0 ¼ ðdobs  dfree Þ=ðdbound  dfree Þ: ð11Þ
curve, but KD is occasionally reported derived from the
(shorter) single parameter observation, see below. Assuming single site binding and fast exchange, it can be
shown that:
3.1. 1D 1H NMR studies of hevein domains 1=X PðboundÞ ¼ 1 þ K D f1=ð½L0  X PðboundÞ ½P0 Þg: ð12Þ

The group of Jimenéz-Barbero has made extensive use of Thus a plot of 1/XP(bound) versus ([L]0 – XP(bound)[P]0) is linear
protein observed chemical shift changes to quantify the bind- with slope KD and intercept 1. Since [P]0 was not accurately
ing affinity of carbohydrates to hevein domains [17]. For the known, it was considered an adjustable parameter and was
most part these studies exemplify the use of non-linear least found by iteration to minimize the linear correlation coeffi-
squares fitting of the observed 1H chemical shifts versus [L]0 cient of the plot. This is not a conventional data treatment
binding curve to find values for KD and dbound. and does not correspond with any of the treatments in Box
Hevein is a small, cystein rich, single chain protein of 43 1. The aromatic signals were selected for study because they
amino acids. The signals from Gln20 NH and Hc in the could often be detected as resolved signals in these modest
1D 500 MHz 1H NMR spectrum can easily be followed dur- molecular weight proteins, and they were also sensitive to
ing titrations with chitobiose and chitotriose [18], or a variety the binding event. This later observation is indicative that
of N-acetyl glucosamine containing oligosaccharides [19], in these residues are in the vicinity of the ligand binding pocket.
order to determine the association constants of the carbohy- The 1H chemical shift changes induced by ligand binding are
drate–protein complexes. In a typical experiment [P]0 was small. Typically, they are within the range 0.01–0.1 ppm and
around 0.7–1 mM and [L]0 reached 15 or 30 mM at the end mostly only around 0.02–0.05 ppm. The values of KD were
of the titration. The experiment protocol was designed so found to be in the range 5 lM to 1 mM.
that [P]0 remains constant as [L]0 varies. This simplifies the
data analysis. A further series of communications from this 3.3. Other examples of 1H detected measurements
group describe the thermodynamics of binding as deter-
mined by variable temperature 1H NMR studies and isother- Trp–Trp (WW) domains are compact modules of 38–40
mal titration calorimetry [20–22]. The close agreement amino acids, folded into a three stranded b sheet. They are
between the NMR derived values and the directly measured found in single or tandem repeats in over 25 unrelated cell
data is a verification of the applicability of the equilibrium signalling proteins. The binding of two phosphothreonine
NMR method to systems with mM binding affinity. peptides to a synthetic construct of the N-terminal WW
domain of Pin 1 has been studied by 600 MHz 1H NMR
titrations [29]. Addition of increasing amounts of peptide
3.2. 1D 1H NMR studies of kringle fragments
ligands caused chemical shift changes for several protons
in the WW domain. During the titration several proton res-
A sequence of papers from the group of Llinás exem-
onances broadened until they disappeared and then reap-
plify the 1D 1H NMR chemical shift titration method
peared at large excess of ligand, indicating slow
and illustrate a special data analysis for the case where
exchange, with the difference in chemical shift between
Dmax can be observed directly, thus in principle reducing
the free and bound forms being of the same order of mag-
the dimensionality of the problem to one. But there was
nitude as the exchange rate. The Ser11 amide proton reso-
some uncertainty in measuring the protein concentration
nance which moved gradually and was more evidently in
so [P]0 was treated as an unknown, thus increasing the
fast exchange throughout the titration was chosen as an
dimensions of the problem back to two.
appropriate marker of the equilibrium conditions. Titra-
Human plasminogen is a single chain protein of 791
tions with 1 mM protein, taken to [L]0/[P]0 ratios of 4.5
amino acids. The molecule has a mosaic structure which
and 11, established values of KD of 117 and 230 lM for
includes five kringle modules, each containing between 78
the peptides by fitting the data to the equation
and 80 amino acids, and with a value of Mr of 9 kDa.
The binding of small molecule ligands to several of the Dobs ¼ 0:5Dmax f1 þ X þ K D =½P0  ½ð1 þ X þ K D =½P0 Þ2  4X1=2 g
domains was studied by following the 1H NMR (300 or ð13Þ
500 MHz) chemical shifts of hydrogens on aromatic resi-
dues such as His, Trp and Tyr during a titration of the where X is the molar ratio of ligand to protein.
ligand up to ratios of around 15:1. The values of KD were During a study of bepridil binding to cardiac troponin C
extracted from the data by either a graphical method [23– it was noticed that addition of the drug caused well defined
27], or by least squares fitting of the hyperbolic binding and specific changes of chemical shift and linewidth in the
1
curve [28]. In their (unique) graphical method the fraction H NMR spectrum (360 MHz) of the protein [30]. These
of ligand bound protein is obtained from the 1H NMR fre- changes occur throughout the spectrum. The most evident
quency of the reporter proton in free protein (dfree), the changes occur in the aromatic region, in the region corre-
226 L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

sponding to the terminal methyl groups of methionine res- the target protein. Dissociation constants are then obtained
idues, and in the aliphatic methyl region. The chemical by monitoring the chemical shift changes of the backbone
shifts of three of these peaks plotted against [L]0/[P]0 pro- amide as a function of ligand concentration. This method
duced smooth curves which reached limiting values near has been widely adopted. It is easy to implement, 15N
the 1:1 ratio. The binding affinity (20 lM) was then esti- labelled materials are often available, the enhanced disper-
mated by the abbreviated method using the directly sion of 2D spectroscopy allows ready identification of suit-
observed saturated shift as a reference. The concentration able reporter signals (often several of them) and the high
of bound protein can be calculated from the known total sensitivity of gradient selected HSQC experiments allow
protein concentration using the relationship rapid data acquisition. This last point is a critical issue. It
Dobs ¼ ½PLDmax =½P0 ; ð14Þ is not practical to follow a titration strategy if the NMR
detection experiment takes a long time.
and KD can be estimated from Eq. (15) which is based on A study of the binding of three potential ligand peptides
the assumption that the concentration of bound ligand to the SH3 domain from the Src family tyrosine kinase Fyn
equals that of bound protein, exemplifies the experiment. The uniformly 15N labelled pro-
K D ¼ ð½P0  ½PLÞð½L0  ½PLÞ=½PL: ð15Þ tein comprised 69 residues. The peptide ligands comprised
14 residues. Multiple binding curves were determined by
A study of the interaction of DMSO with the FKBP12 pro- titrating each peptide into separate samples of 15N SH3
tein is a good example of the utility of NMR to quantify and acquiring 15N-1H HSQC spectra at as many as 14 differ-
weak binding events [31]. Proteins commonly co-crystallize ent ligand:protein ratios from the range 1:10 to 4:1 [32].
with small molecules and solvents, and crystalline FKBP12 Chemical shift changes in both 1H and 15N dimensions were
was known to be associated with six molecules of DMSO. logged and fitted by non-linear regression analysis to
The NMR study was based on the specific perturbation of qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
a few 1H NMR chemical shifts of the protein (0.3 mM) ðK D þ ½L0 þ ½P0 Þ  ðK D þ ½L0 þ ½P0 Þ2  ð4½P0 ½L0 Þ
during a titration with up to 1.5 M DMSO. The binding Dobs ¼ Dmax :
2½P0
isotherm (Fig. 3) clearly covers a large part of the bound
ð16Þ
mole fraction range (see Section 1.3), and is well fitted by
the 1:1 isotherm for KD = 275 mM, a very weak interac- The measured KD values are 3 mM, 50 and 300 lM.
tion. Using an existing 1H assignment and the known ter- The group of Fesik in particular has pioneered the use of
tiary structure of the protein, it was possible to identify protein observed 15N-1H HSQC as a tool for drug discov-
the single binding site. Note that the plot in Fig. 3 is a clas- ery (SAR by NMR [33]) and has used 15N-1H HSQC spec-
sic example of an NMR titration curve and is identical in troscopy for high throughput determination of values of
form to that of Fig. 1. Dd is directly proportional to KD [34–37]. Data are fitted to a single site binding model,
XP(bound) as [L]0 is proportional to [L]0/[P]0. using a least squares fitting search to find the values of
KD and the chemical shift of the fully saturated protein
3.4. Heteronuclear HSQC detected titrations as described above. In this way the protein binding affini-
ties of collections of several dozen ligands at a time can
In these experiments ligand binding is detected from per- be quantified. Observations of protein 15N-1H HSQC spec-
turbations to the inverse detected 15N-1H HSQC spectra of tra in conjunction with ligand titrations is now an estab-
lished favourite method for determining KD by NMR,
but with variations on the data treatment [38,39].
Sometimes Dmax is assumed to be the Dobs at highest
[L]0. Ligand induced chemical shift changes in the
15
N–1H HSQC spectrum of a recombinant two-domain
fragment of barley lectin revealed well resolved, indepen-
dent responses from the two domains, allowing the simul-
taneous determination of the binding affinities of both
sites [40]. Assignment of the 500 MHz 15N–1H HSQC spec-
trum gave dfree for each residue. Chemical shift changes of
several residues were then monitored during titration of the
ligand and dbound was taken from the observations at the
maximum ligand concentration (Mr 9 kDa; [P]0 1 mM;
[L]0–12 mM). This was justified because almost no change
Fig. 3. Chemical shift changes (Dd) for one of the CcH3 resonances of Val- in chemical shifts was observed between the last and the
55 of FKB12 protein as a function of DMSO concentration (correspond-
penultimate titration point, indicating a saturation of bind-
ing to [L]0). The curve represents the best fit solution of the quadratic
equation that describes 1:1 complex formation. The curve corresponds to ing sites. These limiting values were then used to translate
KD = 275 mM and Ddmax = 0.35 ppm [31]. Reproduced with permission. the chemical shift data at each titration point into the frac-
 1999 Kluwer Academic Publishers. tions of occupied protein binding sites (fB and fC) at each
L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242 227

ligand concentration. The concentration of free ligand in upper limit to the size of the protein that can be studied.
the sample is given by There is also no need to isotopically enrich the protein.
Most of the ‘‘traditional’’ ligand observed NMR studies
½L ¼ ½L0  ½PLB   ½PLC ; ð17Þ
have been analysed by linear methods. This is appropriate
and the concentration of protein that remains unbound at because the requirement to observe any high resolution
each domain is ligand signal, that it must be in fast exchange with the
bound form, and present in considerable excess to the pro-
½PB  ¼ ½P0  ½PLB  and ½PD  ¼ ½P0  ½PLD : ð18Þ
tein concentration, is the same as that required to apply lin-
The equilibrium constant characterizing ligand binding to earized data analysis. Hence [L]0  [P]0, [L] = [L]0 and the
any one domain is then given by linearized form of the binding equation is completely justi-
(   fied. The method that has been most widely used is that
1  ½L0 1 based on the relationship
fB ¼  fC  1  
2 ½P0 K D ½P0
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ½L0 ¼ ½P0 Dmax =Dobs  K D ; ð20Þ
 2  ffi)
½L 1 ½L
 f C  1  0  4 0
 f C ;ð19Þ
½P0 K D ½P0 ½P0 and in this case, a plot of [L]0 versus 1/Dobs gives KD as
the intercept on the [L]0 axis.
where f C is the average of the individual values of fC for This is a y-reciprocal plot in Connors’ terms because the
each monitored residue at domain C. dependent variable is plotted on the reciprocal scale. These
A natural continuation from the 15N–1H HSQC protein plots are most frequently presented with [L]0 on the ordi-
detection approach is to apply 13C-1H HSQC detection. nate (y axis) and 1/Dobs on the abscissa (x axis), which is
Hajduk et al. have studied ligand binding to isotopically unfortunate because this orientation obscures any relation-
13
C(methyl) labelled domain 3A of human serum albumin ship with other graphical methods. It would be preferable if
[41]. More than 1800 ligands were screened and binding such graphs were rotated and inverted, so that [L]0 is plot-
was detected by acquiring 13C-1H HSQC spectra ted along the abscissa. An alternative form which is also
(500 MHz) on 50 lM protein solution in the presence and encountered (usually in the 1/T1 studies) has the protein
absence of added compound. The compounds were concentration factored into the y axis
assigned a score based on the magnitude of the change in
the HSQC spectrum and the highest scoring compounds ½P0 =Dobs ¼ ð½L0 þ K D Þ=Dmax : ð21Þ
were progressed to a more complete characterisation by
means of additional ligand titrations. In this way 232 For this solution a plot of [P]0/Dobs versus [L]0 gives KD
ligands with binding affinities in the range 10 lM to when the line is extrapolated to the x axis crossing. This
2.0 mM were measured. form is preferred because the graph is shown the right
way round and the slope gives directly, the bound NMR
4. Ligand observed methods parameter. Whichever plot is chosen, and whichever axis
is chosen to be plotted on the abscissa, all plots give
The drive to make NMR into a tool for drug discovery KD at the intersection with the [L]0 axis, independent of
has lead to a resurgence of interest in ligand detected inter- [P]0.
action studies [42–45]. There has been particular interest in It is worth noting again that in order to apply this kind
the transfer NOE experiment and many new ideas have of data analysis one of the species present has to be main-
been based on intermolecular magnetization transfer. tained at a constant concentration during the course of the
Although primarily designed to detect ligand binding, these titration. Usually this is the species which is not observed in
experiments can also provide information on KD. the NMR spectrum (protein in this section), but this is not
In these experiments the observed species is the small always the case and there are examples of titrations of pro-
molecule and it is almost always titrated into the protein tein into constant concentration solutions of ligand. In this
until present in a considerable excess concentration. It is ‘inverse’ protocol, it is still a requirement that [L]0 at all
convenient to consider the experiments as being of two times greatly exceeds [P]0. This requirement to control solu-
types. In one class, conventional NMR parameters such tion compositions has some implications for the design of
as chemical shift, linewidths, relaxation rates which report the experiment and may (due to solubility, stability and
on the equilibrium condition of the solution are measured. availability) be difficult to achieve.
These data are then processed in a similar fashion to the Two effects might interfere with this kind of study: fail-
preceding section after making appropriate adjustments ure to meet the fast exchange condition, and non specific
to the data analysis to account for the switch of observed binding. Feeney et al. have made a study of the effects of
species. The other class of experiments are based on obser- intermediate exchange processes on NMR observed bind-
vations of intermolecular transfer of magnetization, and ing curves [46]. For a nucleus on a ligand undergoing fast
these have become very popular recently. A major advan- chemical exchange between two sites, the transverse relax-
tage of monitoring the small molecule is that there is no ation rate (and hence linewidth) is given by
228 L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

1=T 2;obs ¼ X PL =T 2;bound þ X L =T 2;free þ X PL X 2L sB 4p2 Dd2 ; site. Several similar ligands were studied at concentrations
ð22Þ around 0.5–12 mM, and in the presence of typically 20–
40 lM protein. The 1H chemical shift effects observed were
where sB is the lifetime of the bound state (equivalent to typically of the order of 0.02 ppm at the highest [P]0/[L]0
1/koff). The last term is the exchange contribution, which ratios and values of KD in the range 2–10 mM were accu-
reduces to zero in the case of very fast exchange. The com- rately measured for several dozen ligand–protein systems.
mon assumption (implied throughout this review) is that Data were analysed by the above mentioned linearization
the protein–ligand complex is in fast exchange on the routine. Under the conditions where the fraction of ligand
NMR time scale, so that whatever NMR parameter is ob- bound to protein is small, KD is given by Eq. (20) and a
served, it is proportional to the mole fraction weighted plot of [L]0 versus 1/Dobs gives KD at the y intercept. The
average of the bound and free states. For many such com- result is shown in Fig. 4.
plexes the association rate constant will be diffusion lim- The N-acetyl methyl group exhibits an upfield shift in all
ited, with a value in the range 107–108 M1 s1. If the of the ligands that bind. The 1H NMR shift of the bound
ligand binds tightly (KD < 0.1 lM) the slow exchange con- ligand is available from the extrapolation of the graph
dition usually applies and separate spectra are observed for shown and the upfield shift for this proton on all of the
the free and bound ligand. For ligands with weaker affinity ligands is around 2 ppm. This is consistent with a model
(KD > 1 mM) the fast exchange usually applies and there where the methyl protons sit directly over the six mem-
are no problems. Feeney et al. point out that in order to bered ring of tryptophan 153.
facilitate the analysis, there is often a general assumption A study of sialyloligosaccharides binding to wheat germ
of fast exchange, without any check that this is in fact agglutinin dimers used the constant [P]0 (0.1 mM), [L]0
the case. Considerable errors (up to two orders of magni- titration (0–15 mM) method again, but extended the data
tude) in KD can arise by indiscriminate assumption of the analysis to include consideration of the ligand linewidths
fast exchange condition. and also incorporated the Swift-Connick equations [50]
In a situation where the ligand is present in considerable as an additional means to estimate the NMR parameters
excess over the receptor it is likely that after saturation of (shift and linewidth) of the bound ligand [51].
the specific binding site, the ligand will bind at non specific The binding of L-tryptophan to the trp repressor pro-
sites. Hence, when interpreting the spectra it should always vides an example of a system where the ligand is not in fast
be borne in mind that the spectra represent an average exchange between the free and bound states. The trp
across all the states that the nuclei experience. It may not
be known a priori how many states there are, or what their
parameters may be. This is a common and recognised
problem in studies of serum albumin (discussed in depth
later), and studies of whole biological entities, e.g., mem-
brane bound receptors. The presence of additional binding
sites has to be acknowledged and built into the model, and
this makes the binding models increasing mathematically
complex. Klotz has provided a very useful account of the
quantitative analysis of this kind of sequential binding [47].

4.1. Ligand observed chemical shifts

One of the early reported applications of this method is


the study by Perkins et al. of the binding of monosaccha-
ride inhibitors to hen egg-white lysozyme [48]. They used
270 MHz 1H NMR to observe the binding induced chemi-
cal shift of the sugar signals. This paper has a useful four
page appendix which derives formulae that are useful for
the analysis of the NMR data from ligand-macromolecule
equilibria, and gives formalisms for a ligand binding in sit-
uations other than the simple two site model.
The binding of sialic acid derivatives to haemagglutinin
was studied by following perturbations to the 500 MHz 1H Fig. 4. Ligand N-acetyl chemical shift data that were used to calculate KD
chemical shift of the sialic acid resonances in the presence for interactions of sialyllactose isomers with haemagglutinins. Each plot
shows the variation of ligand Dd with [L]0 at a constant [P]0. Data for four
of protein [49]. The major perturbation observed was to
different systems are shown. Experiments in panels (a), (b) and (d) were
the chemical shift of the ligand N-acetyl methyl resonance, repeated at different protein concentrations, which is reflected in the fact
presumably due to the proximity of this methyl group to that the second plots have different slopes [49]. Reproduced with
the shielded region of an aromatic residue in the binding permission.  1989 American Chemical Society.
L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242 229

repressor binds two molecules of tryptophan in two inde- 0.23 to 1.1 mM in ligand). In another good example, the
pendent sites with identical affinities. Since no cooperativ- low affinity interaction of antibiotics with bacterial ribo-
ity is involved, the system was treated as a simple two somes were quantified by following ligand 1H transverse
site exchange model. In variable temperature experiments relaxation times, T2 measured by the Carr–Purcell–Mei-
at different protein:ligand ratios the L-tryptophan protons boom-Gill spin echo method [57]. Ligand concentrations
were observed to be in the slow exchange, intermediate were around 0.5–3 mM, ribosomes around 0.2–0.8 lM,
exchange and fast exchange regimes. The low temperature and KD values in the range 0.3–13 mM. The data analysis
data were used to find dfree and dbound for the H-4 proton of is illustrated in Fig. 5.
L-tryptophan. Full line shape analysis of ligand resonances Further discussion of the graphical fitting of linewidth
yielded the dissociation and association rate constants, the data can be found in a report on the binding of sialyloligo-
binding constant, and the thermodynamic parameters of saccharides to wheat germ agglutin, itself a popular subject
the process [52]. of NMR protein–ligand studies [58]. The hits from an
exploratory screening exercise to find new inhibitors of
4.2. Ligand observed relaxation rates human factor Xa were followed by construction of 1H
NMR linewidth isotherms to establish quantitative binding
Because binding induced chemical shift changes are rel- affinities [59].
atively small compared to the linewidth changes, most
ligand observed NMR studies have focused on the relaxa-
tion rate effects. A review of binding induced relaxation 4.2.2. Ligand observed longitudinal relaxation rates (1/T1)
enhancements which includes an excellent discussion of It has been shown that the selective 1H longitudinal
their application to the measurement of equilibrium bind- relaxation rate (1/T1(sel)) of the ligand is a more sensitive
ing constants is that by Ni [53]. indicator of binding than is the nonselective rate [60].
Studies of agonist binding to the acetylcholine receptor
4.2.1. Ligand observed transverse relaxation rates 1/T2 provide straightforward and clear examples of the use of
Fischer and Jardetzky performed the first quantitative T1 data to measure dissociation constants. The relaxation
NMR estimation of a protein–ligand interaction with a times were measured by the inversion recovery pulse
60 MHz 1H NMR study of the interaction of penicillin sequence and the data were analysed by means of the y-
binding to serum albumin, KD = 10–20 mM [54]. These reciprocal plot (Eq.(8)). In a ligand–receptor system where
experiments were performed with penicillin solutions the ligand is present in large excess over the receptor bind-
around 10–500 mM and at a [L]0/[P]0 ratio of around ing sites, the ligand relaxation rate is described by
200:1. At 60 MHz, there was no prospect of making any
1=T 1obs ¼ 1=T 1free þ fbound =ðT 1bound þ sbound Þ; ð23Þ
kind of useful observation of the protein. They estimated
that the maximum likely effect on the ligand chemical
shift upon binding in the vicinity of the diamagnetic
region of an unsaturated ring may be around 200 Hz
(0.33 ppm). Since only about 1% of the ligand was
expected to be bound, the observed change Ddmax for a
fast exchange system was predicted to be about 2 Hz
and this would not be detectable. However the ligand
linewidths were extremely sensitive to the addition of
albumin. This pioneering study demonstrated that the
penicillin–albumin system is indeed in fast exchange,
and that both KD and 1/T2bound could readily be extracted
from the titration data. The calculated relaxation rates (1/
T2bound) for bound penicillin were found to be in the
range 2000–7000 Hz. Shortly afterwards Gerig estimated
the 1H linewidths of tryptophan bound to a-chymotrypsin
at around 30–70 Hz, and put KD in the range 3–12 mM
by observing the ligand line broadening during the course
of a titration [55].
An instructive example is the one by Miller et al. [56]
which describes the binding of choline to the intact mem- Fig. 5. (a) Generic data treatment for the graphical determination of KD
brane bound acetylcholine receptor by measuring the line- from ligand observed NMR data. KD is obtained from the vertical
intercept. The fast exchange parameters Mobs can be Dm, Dd, 1/T2 or 1/T1.
width of the choline methyl 1H NMR (100 MHz) signal
(b) Experimental data from the interaction of the antibiotic telithromycin
during a titration of the ligand. An equilibrium dissocia- (ligand) with bacterial ribosomes. The plots shows [P]0/Dobs versus [L]0 for
tion constant of 190 ± 65 lM was obtained from five data five different reporter protons of the telithromycin molecule [57]. Repro-
points (solutions were 5 lM in receptor and ranged from duced with permission.  2000 The Royal Society of Chemistry.
230 L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

where fbound is the fraction bound ([PL]/[L]0) and sbound is be measured in NMR experiment times that are somewhat
the lifetime of the bound state. The relationship between longer than to those required to measure relaxation rates.
total ligand concentration and T1 is then The reason of course that the diffusion coefficient is use-
ful is that D is closely related to size. A small molecule will
½P0 T 1obs ¼ ð½L0 þ K D ÞðT 1bound þ sbound Þ; ð24Þ diffuse faster than a large molecule. So if the diffusion coef-
ficient of a small molecule is measured in the presence of a
and a plot of [P]0T1obs versus [L]0 gives a straight line with
large molecule to which it binds, the observed diffusion
KD obtained from the intercept of the x axis. This
coefficient will be the weighted average of the coefficients
approach was used to study the binding of various agonists
of the free and bound states. In the presence of a receptor
to the intact acetylcholine receptor of Torpedo californica
protein, the diffusion coefficient of the ligand will be less
Fig. 6, [61]. In an extension of this work, the binding of
than that measured for the free ligand. The usual assump-
a number of cholinergic agonists and antagonists to syn-
tions of fast exchange apply, only this time the ligand needs
thetic and recombinant peptides representative of subunits
to be in fast exchange on both the chemical shift time scale
of the receptor were studied [62]. In a typical experiment,
and also the timescale of the diffusion measurement (usu-
the ligand was titrated into a 0.2–0.3 mM solution of pro-
ally several hundred milliseconds). From the point of view
tein until ligand concentrations around 10 mM were
of data analysis, D is just another NMR parameter, like d
reached. It is not clear that the protein concentration was
or 1/T2 and (Eq. (2)) applies again. Hence,
properly held constant during this protocol.
The serum albumin system also provides several exam- Dobs ¼ X LðfreeÞ DLðfreeÞ þ X LðboundÞ DLðboundÞ : ð25Þ
ples of the use of ligand 1/T1 data to determine KD.
Because of the large amount published on this protein dis- Also, as in the earlier discussion, NMR diffusion studies of
cussion of serum albumins is deferred until Section 4.5. protein ligand interactions are just a special case of more
general intramolecular interaction studies, and again a
large amount of relevant information exists but without
4.3. Ligand observed translational diffusion
the ‘protein–ligand’ label. Two authoritative reviews that
cover the specific protein–ligand area are those by Price
NMR measurements of translational diffusion (D) by
[63] and Lucas and Larive [64].
means of pulsed-field gradient (PFG) spectroscopy have
Larive et al. have coined the term ‘diffusion dynamic
received much attention since the mid 1990s. Essentially
range’ to describe the ability of a PFG diffusion experiment
these experiments use spin echo pulse sequences with a z-
to respond sensitively to the binding constant, and there-
axis magnetic field gradient applied during the first dephas-
fore measure KD accurately [65]. In the context of the
ing time, and again after the refocusing pulse. The effect of
PFG experiment, KD is given by
the first gradient is to spatially encode and the second gra-
dient decodes the nuclear spins. Only spins that are still in K D ¼ ½P0 ðDbound  Dobs Þ=ðDobs  Dfree Þ
their original locations will contribute to the spin echo and
þ ½L0 ðDobs  Dbound Þ=ðDbound  Dfree Þ: ð26Þ
so it is possible, by incrementing the gradient strength or
duration, to measure molecular displacement or translation
diffusion coefficients. Many pulse programs have been
devised to optimize the performance and D can generally

Fig. 6. Plot of nicotine concentration versus [P]0 T1(sel) for multiple Fig. 7. Simulation of the diffusion dynamic range DD (DDfreeDDobs) as a
measurements on two different acetylcholine receptor preparations (circles function of the ligand:protein ratio for various dissociation constants, (A)
and squares). The pyridinyl H-4 proton of nicotine was used for the KD = 1000 lM, (B) KD = 100 lM, (C) KD = 10 lM, (D) KD = 1 lM and
relaxation measurements. Similar results were obtained from the other (E) KD = 0.001 lM. The simulation was performed for a fixed [P]0. The
aromatic protons. The data show the typical scatter in the selective T1 diffusion coefficient of the ligand (DDfree) was set at 6.9 · 1010 m2 s1,
measurements, and estimated error bars [61]. Reproduced with permis- and that of the protein (DDbound) at 0.76 · 1010 m2 s1 [64]. Reproduced
sion.  1988 Biophysical Society. with permission.  2004 Wiley Periodicals, Inc.
L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242 231

Fig. 7 was produced by calculating the values of DD for a 0.39 ± 0.26 mM. This plot shows how the observed diffusion
hypothetical system with ligand:protein molar ratios rang- coefficient of the ligand is strongly attenuated by binding and
ing from 0.1 to 500 and for five different values of KD. The is very sensitive to the ligand:protein ratio at low [L]0
useful dynamic range is region 2 where Dobs depends on (<25 mM). At higher [L]0 most of the ligand is unbound,
both KD and the ligand:protein ratio. Region 1 corre- so the curve levels and approaches Dfree.
sponds to low ligand:protein ratios, Dobs–Dbound and DD Lennon et al. [66] noted that this PFG 31P NMR
converges to the difference between Dfree and Dbound. At method has an advantage over the more obvious 31P chem-
too high ligand:protein ratios (region 3), Dobs–Dfree and ical shift perturbation method because several factors other
DD converges to zero. than binding to haemoglobin influence the 31P chemical
Note that this graph is simply a special case of the dis- shifts. The PFG NMR method is one of only a few tech-
cussion of Section 1.3. It is axiomatic that an experiment niques available in which these interactions can be studied
aimed at measuring KD has to be sensitive to the binding in intact erythrocytes. However the experiments do take a
event. The diffusion experiment is measuring a property long time. Each titration data point is the result of an incre-
of the ligand which changes as a function of the ratio of mented NMR experiment that especially at low ligand con-
the free and bound forms. centrations required substantial signal averaging. The data
A 162 MHz 31P NMR study of the binding of 2,3-biphos- shown in Fig. 8 are clearly in the correct part (region 2) of
phoglycerate to haemoglobin is the first documented appli- Larive’s diffusion dynamic range graph. An earlier study of
cation of the diffusion technique to a protein–ligand the same 2,3-biphosphoglycerate/haemoglobin system adds
complex [66]. The diffusion coefficient of the ligand was mea- a novelty footnote to this discussion [67]. The 1990 study
sured from a series of solutions that were around 3–5 mM was an equilibrium dialysis determination of KD and 13P
haemoglobin and ranging from 2 to 80 mM ligand over 10 NMR was used to determine the concentration of 2,3-
data points (see Fig. 8). The data analysis consisted of fitting diphosphoglycerate with the dialysis cell nestled inside a
Dfree and KD to the titration curve, with Dbound set to the 10 mm NMR tube.
value of Dprotein that was measured in a separate experiment. Some 15N filtered diffusion experiments have been used
The resulting values of KD for the binding of 2,3-biphospho- to study of the binding of a peptide to the Src homology
glycerate with carbonmonoxy-, oxy-, and deoxy- haemoglo- 3 domain of phox47 (60–70 amino acids). A binding iso-
bin are, respectively, 1.98 ± 0.26 mM, 1.8 ± 0.5 mM and therm was constructed from the diffusion coefficients of
the species in solutions that were 0.46 mM in protein and
ranging from 0.23 to 10.5 mM in ligand (ligand:protein
ratios from 0.5:1 to 23:1) and Dobs varied over the fairly
narrow range 0.95 · 1010 m2 s1 to 1.55 · 1010 m2 s1.
This was sufficient to measure KD at 21 ± 14 lM, compara-
ble with a value of 29 ± 3 lM established by fluorescence
spectrophotometry [68].
It is fair to ask ‘‘what is the advantage of diffusion based
estimates of KD over more classical relaxation rate experi-
ments?’’ The preferred experiment will be the one that best
discriminates between the bound and non-bound forms of
the ligand. For a small molecule binding to a large protein
in dilute non viscous solution D is likely to change by
around a factor of 10. This is a good ‘dynamic range’.
The relaxation rate difference between bound and free
states will be much more variable, depending on what
nucleus is observed and what relaxation rate is measured,
but will often be in the same range or greater. Probably
relaxation rate experiments are more direct and faster.
One clear advantage of the diffusion coefficient
approach is the possibility of determining KD from single
data point experiments, thus eliminating the need for
assembling a titration curve at different ligand concentra-
Fig. 8. The diffusion coefficient of 2,3-biphosphoglycerate (DPG) in tions. In contrast to most of the preceding experiments
carbonmonoxyhaemoglobin solutions as a function of [DPG]. Each value where the NMR parameter of the bound state (Dmax) can-
of D was derived from a PFGLED NMR experiment. The three not be directly observed, (the exception is that sometimes
parameters that define the solid line are the diffusion coefficient of non-
Dmax is assumed to have been observed in some protein
bound DPG (Dfree = 1.8 · 1010 m2 s1), the diffusion coefficient of the
carbonmonoxyhaemoglobin (Dbound = 0.1 · 1010 m2 s1), and the disso- detected studies), the two limiting diffusion parameters,
ciation constant (KD = 1.98 mM) [66]. Reproduced with permission.  that of the ligand (Dfree) and that of the protein–ligand
1994 Biophysical Society. complex (Dbound) can both be directly measured in separate
232 L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

PFG experiments. The diffusion coefficient of the bound and free states. Hence there is a risk that part of the bind-
ligand is simply that of the protein. In practise many work- ing curve will originate in a zero signal to noise region.
ers continue to create binding isotherms from titration data
as discussed above, but the short cut route to single point 4.4.1. Saturation transfer difference NMR
measurements is attractive for high throughput determina- The saturation transfer difference (STD) experiment was
tions and has been demonstrated as a viable means to devised to screen compound collections for binding activity
quantitatively rank order screening hits from the SHAPES to proteins [72]. Saturation transfer refers to the mecha-
strategy for drug discovery [69]. The cost of a faster exper- nism whereby magnetization introduced into a large pro-
iment is some degradation in the precision of the method. tein molecule is able to very rapidly spread around all of
Small errors in the measured diffusion coefficients arising the constituent spins, including any attached ligand, such
from any sources will result in large errors in KD. A simple that when the ligand leaves its binding site it carries with
analysis based on propagation of random errors found that it information in the form of spin polarization from the
in the specific case of the SHAPES screen a 5% error in protein. Thus the bulk NMR response of the ligand is an
Dobs translated to a 125% error for KD = 1 mM, a 57% averaged response, composed of the sum of signals from
error for KD = 100 lM, and a 127% error for KD = 10 lM non interacting ligands and previously bound ligands. This
[69]. This is yet another manifestation of the material dis- saturation process is very efficient, so the modulation of the
cussed in Section 1.3. ligand signal induced by the protein is readily detected,
even in the presence of a large excess of ligand. In order
4.4. Ligand observed magnetization transfer to make the interaction apparent, the experiment is imple-
mented as a difference experiment. Two 1D spectra are
Two completely new NMR methods for measuring dis- acquired– one with selective excitation of the protein
sociation constants have been available since 1999-2000. turned on, and one where the selective excitation is moved
Both techniques are based on observing the intensity of to an empty spectral region. The difference spectrum will
magnetization transferred to free ligand via the bound show a response only of ligands that were at some time
ligand from the protein protons. This is accomplished associated with the protein. The sensitivity to discriminate
either by direct selective excitation of the protein protons, between binders and non binders is excellent, and this
or by relaying magnetization via the solvent and protein. experiment has been successfully used to identify ligands
The responses of magnetization transfer experiments are with binding activity from multicomponent mixtures. The
not classical NMR parameters like chemical shifts or relax- STD experiment has been authoritatively reviewed by
ation rates. They are exchange averaged parameters, but Meyer and Peters [44]. A recent review by Krishnan con-
their magnitudes are not simply determined by the respec- tains a good discussion of the quantitative analysis [73].
tive populations and NMR parameters of free and bound How is the STD response related to KD? Because the
states, so Eq. (2) cannot be applied. The responses do STD experiment is a difference experiment, the STD spec-
not have absolute values, they are dimensionless and are trum contains only signals from the bound state of the
affected by a host of experiment parameters. These experi- ligand. Hence the STD response reports on the concentra-
ments are fundamentally different to those discussed in the tion of the protein–ligand complex. As with all of the pro-
preceding sections because they report on the concentra- cedures so far, a titration with ligand is carried out to map
tion of bound ligand [LP]. the response as a function of [L]0. Some normalization of
The well known transfer NOE effect might be regarded the STD signal intensity is then required. A relative STD
as the forerunner to the following two methods. It is the effect is defined by normalizing the STD signal to the inten-
application of the 2D NOE experiment to exchanging sys- sity of the same peak in the off-resonance spectrum, and
tems. Typically a small molecule (ligand) is characterized then a correction for total ligand concentration is intro-
by a short correlation time and small positive NOE values. duced to arrive at an ‘STD amplification factor’ ASTD. A
A large molecule (protein) is characterised by a long correla- plot of the parameter ASTD versus [L]0 is a normal binding
tion time and large negative NOE values. So that for a pro- isotherm and can be fitted to derive KD [74]. So, although
tein–ligand system where the ligand is present in excess the observed STD NMR intensity is not a direct measure of
[L]0 > [P]0 and is in fast chemical exchange, the large negative the affinity of a particular ligand, KD can be obtained from
NOE acquired by the ligand while it was at the binding site the titration curve ASTD versus [L]0. The method is illus-
overwhelms the small positive NOE of the unbound ligand. trated by Fig. 9 which shows the binding of methyl b-D-
The observed response will be dependent on the fraction of galactoside to the 120 kDa lectin ricinus communis agglutin
bound ligand (amongst many other parameters) [70]. I (RCS120) [74].
The potential of NOE experiments to be used quantita- Ligand binding to human integrin aIIbb3 incorporated
tively in this fashion seems not to have been explored into liposomes has also been studied in this way. This mem-
except to note that the affinity of ligands can be rank brane bound fibrinogen receptor consists of two sub units
ordered from the NOE pumping response [71]. A funda- 125 and 108 kDa. The binding affinity of the peptide ligand
mental problem with using the transfer NOE response to cyclo(RGDfV) was estimated at 30–60 lM, from observa-
quantify binding is that it changes sign between the bound tions of a solution that was 5 lM in protein and 29–
L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242 233

[79]. The result is that the resonances of non-binding com-


pounds appear with opposite sign and tend to be weaker
than those of interacting ligands. In order to make quanti-
tative estimates of KD, some normalisation of the Water-
LOGSY signal intensity has to be made. This is necessary
because the ligand acquires some magnetisation directly
from bulk solvent irrespective of what is happening at the
binding interface. The correction is made simply by per-
forming the WaterLOGSY experiment again, this time on
the ligand solution without the protein present. The exper-
iment is performed at several different ligand concentra-
tions and the corrected response can be plotted against
[L]0 to produce the normal binding isotherm. This is anal-
ogous to the correction that has to be made to ASTD in the
previous experiment. The experiment is illustrated in
Fig. 10 with data from the tryptophan/human serum albu-
min system. The circles are the uncorrected responses and
Fig. 9. (a) The normalised saturation transfer difference response of the
the triangles are the response of ligand without protein that
H-4 signal of b-GalOMe as a function of [L]0/[P]0 during a titration of b-
GalOMe into 40 lM RCA120 protein. (b) Display of the same data in are used to apply the correction. The corrected response
terms of the STD amplification factor. This second plot shows that even has the profile of a familiar binding curve.
though the fraction of ligand which is saturated decreases at a higher
ligand concentrations, the absolute STD signal intensity increases in the
form of a saturation curve. The STD amplification factor is obtained by 4.4.3. Comment
multiplying the fractional STD effect shown in (a) with the ligand excess The magnetization transfer experiments are best viewed
[74]. Reproduced with permission.  2001 American Chemical Society. as a measure of the bound ligand concentration, for a sit-
uation where the proportionality constant between the
spectroscopic response and [PL] is unknown. In this respect
275 lM in the peptide, corresponding to ligand:protein
the data are not different to, for instance, the chemical shift
ratios from 6:1 to 55:1 [75]. A more recent STD NMR
titration where the limiting chemical shift is unknown. So
study reports on the binding of a peptidomimetic ligand
there is no difficulty in producing binding curves such as
to human CD4 protein [76]. STD amplification factor titra-
Figs. 9 and 10 and fitting them to a parabolic curve. KD
tion curves from several of the ligand protons were pre-
arises from the curvature, thus acquiring the concentration
sented. The STD derived KD (9 lM) compares favourably
with that derived from surface plasmon resonance (10 lM).
The STD method is greatly advantaged for ligand–pro-
tein interaction studies because it is a method that is com-
pletely unlimited by the size of the protein. It works better
for larger proteins. Most of the published STD NMR stud-
ies that have been published in the last five years are aimed
primarily at gaining structural information by epitope
mapping. These reports have not usually addressed the
issue of measuring dissociation constants, other than two
studies which rank-ordered the ligands according to their
(unquantified) binding affinities [77,78].

4.4.2. WaterLOGSY
In the second magnetization transfer experiment, bulk
water magnetization is transferred to the ligand via the
ligand–protein complex. This is termed WaterLOGSY
(Water-Ligand Observed by Gradient SpectroscopY).
The experiment relies upon the water molecules present Fig. 10. WaterLOGSY responses for the C2-H resonance of L-tryptophan
at the protein–ligand interface and uses intermolecular as a function of ligand concentration during a titration into protein. The
NOE and chemical exchange with labile hydrogens to circles and triangles are the experimental intensities recorded in the
transfer magnetisation from bulk water to the protein. presence and absence of 10 lM HSA, respectively. The WaterLOGSY
response without protein is linear with [L]0 and can be used to apply a
The acquired magnetization is in turn transferred to any
correction for effect of non bound ligand. The square points are the
bound ligand which, when appropriate dissociation rates intensity difference graph and the line is the best fit calculated quadratic
apply, can leave the binding site carrying with it magnetisa- curve. The signal intensity is on an arbitrary scale [79]. Reproduced with
tion which has the same sign as the starting magnetization permission.  2001 Kluwer Academic Publishers.
234 L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

units of [L]0, and the proportionality constant is given by Table 1


the asymptote, but is a meaningless parameter. Dissociation constants (KD (mM)) and stoichiometries (n) obtained from
NMR studies of ligand binding to serum albumins
Neither the direct STD experiment, nor the direct
WaterLOGSY responses appear to have been widely used Protein Ligand NMR KD n Reference
method (mM)
for the determination of KD. The disadvantage of these
methods appears to be the need to correct the response HSA TFBAa D 2.2 9 [81]
BSA Phenylbutazone 1/T2 3 100e [82]
of these experiments with a second reference experiment BSA Azathioprine 1/T2 3 100e [83]
which doubles the length of what is already a lengthy pro- HSA Ibuprofen D 17 ± 4b 50 [84]
tocol. For a simple system consisting of only one protein HSA Ibuprofen 1/T1 49 ± 15b 36 ± 8b [84]
and one ligand, chemical shift or relaxation rate methods HSA Ibuprofen 1/T1 1.5 ± 0.2 22 ± 1c [85]
will be faster and easier to implement. The STD and HSA Ibuprofen 1/T1 2.0 ± 0.4 15 ± 2d [85]
BSA Ibuprofen 1/T2 50–100 3–7 [86]
WaterLOGSY experiments would be expected to work HSA Ibuprofen 1/T1 1.4 ± 0.2 33 ± 2 [87]
and would have an advantage when directly quantifying HSA Salicylate D 17 ± 8 35 [88]
binding in a mixture of tentative ligands. Both STD and HSA Salicylate 1/T1 15 ± 2 32 [88]
WaterLOGSY have been used to monitor binding (quanti- BSA Salicylate D 30 ± 4 33 ± 3 [89]
tatively) in competition binding studies and these studies BSA Salicylate Dd 1.2 ± 0.6 8±3 [86]
HSA Salicylate 1/T1 4.3 ± 0.5 35 ± 2 [87]
are discussed in Section 6.
a
4-Trifluoromethylbenzoic acid.
b
Entry is the simple mean and standard deviation of the range of results
4.5. Ligand observed binding to serum albumins quoted in the original communication.
c
pH 6.8.
Ligand detected studies of protein–ligand systems are d
pH 8.0.
e
vulnerable to interference from non-specific binding. Math- Assumed.
ematical models of multisite binding have been summa-
rized by Klotz [47]. The simplest approach, and the one
that is widely used, is to assume that the protein has n few) molecules that bind specifically make no significant
equivalent and non-interacting binding sites, contribution to the experimental observable which is dom-
inated by the behaviour of non-specific binding interac-
P þ nL ¡PLn
tions and free ligand.
and therefore The PFG diffusion method, relaxation rate methods and
n chemical shift perturbation methods have all been success-
K D ¼ ½P½L =½PLn : ð27Þ
fully applied in several studies of non-specific binding to
In this model the parameter n works as a scaling on the serum albumins. Table 1 summarises some of these studies.
NMR parameter of the bound state (Dmax). The bound This table illustrates the wide range of ligand observed
population of ligand given by NMR parameters that can successfully be used as a handle
1=2 on the dissociation constant. Also noteworthy are the high
X LðboundÞ ¼ a  ða2  bÞ ; ð28Þ
values of n showing that as well as being a promiscuous
where binder of ligands, serum albumins are also polygamous.
Also noteworthy are the wide range of reported values of
a ¼ ð½L0 þ n½P0 þ K D Þ=2½L0 ; ð29Þ
KD, indicating that either these data are not precise mea-
and sures, or that the system is itself is somewhat fuzzy. Both
explanations are valid. The discussion in the later salcy-
b ¼ n½P0 =½L0 : ð30Þ
late/BSA report contains a more complete account of
Eq. (28) can be inserted into Eq. (2) to calculate a binding non-specific drug-albumin interactions and NMR studies
curve to match experimental NMR data. The widespread [86].
use of this formula is undoubtedly due its simplicity rather It should be recognised that for systems with multiple
than its validity. It is straightforward to introduce the new binding sites, and where the stoichiometry (n) is introduced
term n into computer enabled data fitting routines. as an additional parameter, the binding curve is now a four
Serum albumins are able to bind drugs specifically at parameter fit (Mfree, Mbound, KD and n) and a cosmetically
high affinity binding sites (KD = lM), and non-specifically appealing fit is guaranteed to be obtained for the limited
at several, possibly dozens, of low affinity sites (KD = mM) number of data points that define the usual NMR experi-
[80]. It commonly occurs that different studies of drug- ment. Further, the four parameters are highly interdepen-
serum albumin interactions report binding affinities that dent and it may not be possible to obtain a unique
differ by orders of magnitude because different protocols solution. As always, a good fit of the calculated curve to
explore different parts of the specific–non-specific binding the experimental data is not a validation of the binding
continuum. Ligand observed NMR methods are carried model. A good fit of a four parameter curve to a handful
out at high ligand:protein ratios and so by definition are of data points should justly be viewed with some
weighted to report on non-specific binding. The one (or a scepticism.
L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242 235

5. Competition binding experiments

A general drawback of all ligand observed NMR meth-


ods is their failure to work directly with high affinity
ligands. It is generally accepted that the lower limit of
applicability of the NMR method is KD to 10–100 lM.
Ligands with dissociation constants lower than this are
tightly bound which means that there is little exchange with
free ligand during the time scale of the NMR experiment
(typically a few hundred milliseconds) and therefore no
information about the bound state is detectable. This limit
has become smaller as the sensitivity of NMR hardware
has increased. This difficulty can be overcome by exploiting
the well known phenomenon of competition binding. The
binding experiment is performed in the presence of a sec-
ond ligand which occupies the same binding site as the tar-
get ligand. The competition process effectively modulates
the NMR response of the observed ligand and scales it
back into the region p = 0.2–0.8. No new NMR detection Fig. 11. 500 MHz 1H NMR linewidth data from the titration of a(2,6)-
sialyllactose (L1) using Neu4Pp5Aca2Me (L2) as a reporter ligand to
techniques are introduced in the following discussion.
probe the interaction with haemagglutinin. [L2]0 is fixed at 1.0 mM and
The experiments are based on exactly the same parameters [P]0 is fixed at 2 mg/ml. (a) The linewidths of two protons (N-acetyl methyl
(e.g., linewidths, relaxation rates, and NOE responses) that and 2-O-methyl) on the reporter ligand (L2) as a function of [L1]0. (b) The
have been previously described in Sections 3 and 4, except response of the N-acetyl methyl protons of the analyte ligand (L1) as a
that they are now applied to a three component system. function of [L1]0. In both cases KD is derived from the y- intercept [90].
Reproduced with permission.  1992 Academic Press, Inc.
The experiment protocols and the data analysis methods
are the points of interest.
A few of the competition methods have been purpose
designed to push the limits of NMR quantification to L1 or L2 is the observed species. In the case where the con-
sub-micromolar levels. Other studies have arisen from centration of the measured ligand [L1]0 is varied while the
NMR screening methods to identify new ligands for phar- NMR property of the reference ligand L2 is observed
macologically interesting receptors and were driven by a  
K D DmaxðL2Þ ½P0 ½L20
need to circumvent the false negative issue with high affin- ½L10 ¼  KD 1 þ ; ð31Þ
ity ligands. Although designed primarily to detect and sig- K DðL2Þ DobsðL2Þ K DðL2Þ
nal the binding event, it turns out that these data can also and a plot of [L1]0 versus 1/Dobs(L2) is a straight line with a
be used, often quickly and with little extra effort, to quan- y-intercept of –KD(1 + [L2]0/KD(L2). Thus, if KD(L2) has
tify KD. been measured independently, KD can be determined from
the chemical shift or (as in this case) from the linewidth
5.1. Graphical data treatment behaviour of species L2 as a function of [L1]0. This proto-
col is useful for ligands whose resonances are difficult to
A study of the binding of some sialic acid derivatives to analyse due to overlap or poor signal-to-noise ratio. The
haemagglutinin in intact influenza virus is the first reported outcome is illustrated in Fig. 11a from which KD =
application of quantitative NMR competition methods to 2.7 ± 0.2 mM.
the study of protein–ligand interactions [90]. Although The other scenario is where the NMR parameter of the
nowadays this method is not likely to be used, this commu- titrating ligand L1 is monitored as a function of varying
nication is instructive because it contains clear accounts of [L1]0. In this case it can be shown that
the data analysis. The linewidth of the N-acetyl or O-
methyl 1H resonances of a reporter or reference ligand  
DmaxðL1Þ ½P0 ½L20
(L2) were followed during titration with sialic acid deriva- ½L10 ¼  KD 1 þ ; ð32Þ
tives of interest (L1). Typically titrations were carried out DobsðL1Þ K DðL2Þ
at 1 mM in the reference ligand and the ligand of interest
was incremented up to 15 mM. As is usual in graphical and a plot of [L1]0 versus 1/Dobs(L1) is again a straight line
methods the solution compositions were contrived to bring with a y-intercept of –KD(1 + [L2]0/KD(L2). This is illus-
about a reduction in the number of degrees of freedom of trated for the same ligand system in Fig. 11b, which leads
the system. In this case the titrations were performed at to KD = 1.8 ± 0.5 mM.
constant reference ligand concentration [L2]0 and constant In these experiments the reporter ligand has a binding
protein concentration [P]0. Graphical data analyses were affinity which is not very different from those of the inter-
presented for two possible cases depending on whether esting ligands. It is an important study because it expands
236 L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

the applicability of the NMR titration method to systems


where tighter binding constants may be determined. Eqs.
(31) and (32) are based on an assumption that the concen-
tration of free ligand is approximately the same as [L]0, so
the dissociation constants that can be measured are limited
to the range [P]0.

5.2. Magnetization transfer for sub micromolar binding

Some more recent developments to competition binding


have successfully extended the range of NMR measurable
values of KD to lM levels. The interest in this topic can
be traced back to a communication from Mayer and Meyer
[74]. They observed that the STD response of a bound
ligand was reduced by the presence of a competitor ligand,
and noted that it is possible to determine the KD value of a
ligand from the IC50 value of any competitor ligand which
has a known dissociation constant, thus
Fig. 12. WaterLOGSY signal attenuation of the reference compound as a
K D ¼ ð½L10 K I Þ=ðIC50ðL2Þ  K I Þ: ð33Þ function of the dissociation constant KI of the competitor. The simulation
was performed using Eq. (34) with a competitor concentration of 5 lM, a
In the above equation, and in the remainder of this section, reference (ligand of interest) concentration of 50 lM, and a protein
the symbol KI is used to indicate the dissociation constant concentration of 2 lM [91]. Reproduced with permission.  2002 American
of the competitor, reference or reporter ligand (L2) and KD Chemical Society.
is used to indicate the dissociation constant of the ligand of
interests (L1). served ligand and the ratio I(+)/I() approaches zero. The
Dalvit [79,91] recognised that the waterLOGSY four curves illustrate the dependence of the function on
response (see Section 4.4.2) of a system comprising of a the binding affinity of the reference molecule. Weaker affin-
receptor and a specific, but medium affinity (mM) ligand ity reporter ligands are more easily displaced. The effect of
could be used as a screen to detect the presence of compet- higher affinity reference molecules is to move the curves to-
ing higher affinity ligands. Remembering that the water- wards the left and top of the graph (mirroring the effect of
LOGSY response arises from bound ligand, any molecule decreasing KD).
that competes with the bound ligand will result in a The experiment has been validated with the model sys-
decrease in intensity of the waterLOGSY signal. Knowl- tem serum albumin with 6-methyltryptophan as a reference
edge of the binding affinity of the reporter ligand, coupled ligand (KD = 37 lM) and diazepam as the high affinity
with some careful experiment design permits quantitative competitor ligand. A single point experiment established
determination of the binding constant for the competition a 65% reduction in the reporter ligand waterLOGSY signal
ligand. Setting up the experiment conditions (solution com- when diazepam was added. This translated to
positions) so that waterLOGSY responses of investigative KD = 2 ± 1 lM for the diazepam binding [91]. The same
ligand (L1), reporter ligand (L2) and protein are compared group have demonstrated that transverse relaxation and
with directly corresponding reporter ligand (L2) and pro- longitudinal relaxation parameters of a reference ‘reporter’
tein data simplifies the data analysis. The relationship ligand are equally effective for quantifying high affinity
between attenuation of waterLOGSY response I(+)/I() binding [92]. Further work on competition methods from
and the dissociation constants of the two ligands is given this group are discussed in Section 6.3.
by The HSA/tryptophan/diazepam system was also used
  rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
n  o2 recently to demonstrate that competition STD NMR can
½L20 ½L2
I ðþÞ ½P0 þ ½L10 þ K D 1 þ KI  ½P0 þ ½L10 þ K D 1 þ K I 0  4½P0 ½L10 also be put into a quantitative context [93]. The principles
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2
I ðÞ
½P0 þ ½L10 þ K D  ½P0 þ ½L10 þ K D  4½P0 ½L10 and experiment protocols are almost identical to that
ð34Þ
described above, except that magnetization transfer is
induced by the STD pulse sequence instead of water-
Fig. 12 is plotted directly from solutions of Eq. (34) to illus- LOGSY. Again, binding of the high affinity ligand is sig-
trate the dependence of the response ratio I(+)/I() on [P]0, nalled by a decrease in the intensity of the reporter ligand
[L1]0, [L2]0, KD and KI, respectively [91]. This chart demon- STD response, and the magnitude of this attenuation is
strates that the ratio I(+)/I() is effectively modulated as a used to deduce KI for the strong binder. The dissociation
function of the inhibitor binding affinity. The inhibitor dis- constant of diazepam was estimated at 2.4 ± 0.5 lM from
places the monitored ligand, thus reducing the water- the observed fractional reduction of 0.41 in the 6-methyl-
LOGSY response. The chart shows that more effective tryptophan STD signal. Fig. 13 is complementary to
inhibitors (KI approaches zero) displace more of the ob- Fig. 12. It shows the modulation of the STD response of
L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242 237

deconvolute the bound and free mole fractions of the spe-


cies that is observed and the only change is that we are now
looking at 19F as a reporter nucleus. This review could well
have been written with the discussion of 19F experiments
sprinkled around the text under the other appropriate
sub headings. However there is a didactic advantage in pre-
senting this material separately. At the risk of being repet-
itive, the papers cited in the following discussion illustrate
all of the experiment protocols and data analysis tech-
niques again, but often with greater clarity.

6.1. Fluorine labelled proteins

A study of oligosaccharide binding to AcAMP2-like


Fig. 13. STD signal reduction of the indicator as a function of the peptides provides a recent example. A 30 amino acid
competitor concentration. The concentration of the STD indicator residue analogue corresponding to the hevein domain was
(KD = 10 lM) and the protein are 100 and 5 lM, respectively. Simulations synthesised with Phe18 and Tyr20 changed to 4-fluorophe-
were performed according to Eq. (20) for five different dissociation nylalanine. The mM binding of chitotriose was evaluated
constants (KI) of the inhibitor (100, 10, 1, 0.1 and 0.01 lM). The STD
from 19F NMR titration data [97].
signal reduction is expressed as a fraction and is equal to the ratio of the
STD signal intensities in the presence and absence of the competitor. [P]0 is
included in the simulation. The dashed lines represent simulations with [P]0 6.2. Fluorine labelled ligands
omitted [93]. Reproduced with permission.  2004 John Wiley & Sons, Ltd.
The binding of N-trifluoroacetyl-D-(and L-)p-fluorophe-
a reporter ligand as a function of the inhibitor concentra- nylalanine to a-chymotrypsin was quantitatively measured
tion for a range of values of KI. This chart shows that from perturbations to the 19F chemical shifts of both fluo-
the fractional reduction of the STD signal of the reporter rine nuclei upon interaction with the active site [98]. The
is most sensitive to the competitor concentration in the downfield shifts of the bound ligand are attributed to inter-
range 0.2–0.8 (this is yet another statement of the material actions with His, Ser and Asp residues in the binding
in Section 1.3). The chart suggests that there is no lower pocket. The communication includes a full and clear
limit to the KI values that can be measured in this way. description of a unique data analysis which is based on
The KI of any higher affinity ligand can be accurately mea- arithmetic iterations of the quadratic binding equation to
sured by lowering the concentration of inhibitor until the determine Dmax and KD. Although obsolete, the procedure
STD modulation is in the 0.2–0.8 range [93]. is worth noting. Mathematical procedures to account for
Another discussion of NMR detected competition bind- dimerization of the protein add interest to the data analy-
ing noted that KI was in principle available from knowl- sis. Data were collected under conditions of constant [P]0
edge of the KD of the reporter ligand but stopped short (1.9 mM) and varying [L]0 (from 0.5 to 20 mM); and for
of reporting quantitative figures [94]. the reverse case of constant [L]0 (1.5 mM) and varying
[P]0 (0.05 to 1.9 mM). It was found that the L isomer binds
19
6. F NMR studies about 10 times less strongly than the D isomer
(KD = 0.32 mM at pH 6.0).
Introduction of a fluorine atom into either the protein or A consequence of the large chemical shift range of 19F
the ligand opens the door to 19F NMR observation and is that dynamic systems may often be in the intermediate
this has an enormous range of benefits. 19F has nearly or slow exchange regime when observed via 19F. In the
the same sensitivity as 1H. It is a spin 1/2 nucleus and its slow exchange case the data treatment becomes trivial.
chemical shift range in proteins and small organics is usu- Integration of the free and bound signals gives [L] and
ally around 100 ppm. There is essentially no biological [LP] directly. Much of the 19F NMR protein–ligand
background of 19F signals, thus only the site of interest will interaction literature reports on systems that are on this
appear in the experimental spectrum. These factors make time scale. A study of the binding of trifluoromethyl ana-
the 19F NMR spectrum easy to acquire and easy to analyse. logs of 6,7-dimethyl-8-ribityllumazine to lumazine apo-
Furthermore, 19F occurs frequently as a motif in man made protein is such an example. The bound ligands give
drugs (it is usually put in to improve adverse metabolic rise to additional broad signals several ppm downfield
profiles). Consequently there is a large body of literature of the free ligand signal, KD = 3–5 lM, analysed by Scat-
that reports on the use of 19F NMR to study protein– chard plots [99]. This system is discussed further in Sec-
ligand interactions [95,96]. tion 7.
19
From the perspective of quantifying the protein–ligand F spin–spin relaxation rates (1/T2) measured by the
interaction, nothing that follows is new to what has been CPMG sequence have been successfully applied to measure
presented in the preceding sections. The concern is still to the low affinity binding of the volatile anaesthetic isoflu-
238 L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

rane to BSA [100]. In this study, only one 19F signal was ysis of the ligand H-4 proton over a range of protein and
observed, but exchange is sufficiently slow that exchange ligand concentration (mM) and over a temperature range
broadening had to be allowed for [54]. This was achieved from 20 C where the system is in slow exchange to 65 C
by varying the interval between the 180 refocusing pulses. when it is in fast exchange. This thorough analysis results
The data analysis then proceeded exactly as described pre- in a full thermodynamic picture of complex formation in
viously in Section 4.2. A plot of 1/T2(obs) versus [L]0 gives – the form of a KD versus 1/T Arrhenius plot.
KD at the x intercept, (1.4 mM). The thermodynamic and kinetic parameters associated
with the binding of N-acetylgalactosamine to Artocampus
6.3. Fluorine observed competition binding experiments integifolia agglutin were determined from the temperature
dependence of line broadening in the 19F and 13C NMR
The large chemical shift anisotropy of 19F results in very spectra of the ligand [104]. In this system chemical shift
broad lines for bound fluorinated ligands and therefore changes were not observed on binding. Campbell et al.
very large differences in line widths between the bound reported a full line shape analysis of the 1D profiles of indi-
and free states. This makes fluorinated ligands particularly vidual 15N–1H HSQC peaks at each point of a titration of a
well suited for competition binding experiments. The trans- 12-residue phosphopeptide into a solution of the SH2-N
verse relaxation rate 1/T2 of fluorine in a reference ligand is domain of the p85a subunit of PI 3 0 -kinase [105]. This anal-
an ideal parameter to monitor as a function of test ligand ysis gave information about the kinetics of complex forma-
concentration. For instance, the low affinity ligand 2- tion in a system with KD in the nM range. A software tool
hydroxy-3-fluorobenzoic acid has been used as a reporter is available to facilitate the analysis of line shapes from
ligand for the Sudlow site 1 of human serum albumin. Dis- titration generated two-dimensional spectra [106].
sociation constants for hundreds of compounds in the Again, 19F NMR observations of fluorinated ligands
range from a few nM to high lM are claimed to have been provide some of the clearest examples of such studies. All
measured by this method [101]. of the kinetic parameters, including kon and KD were
Dalvit et al. have recently provided an in depth theoret- obtained from quantitative analysis of NOESY spectra of
ical analysis of 19F NMR competition binding using weak the lumazine protein/ligand system [107]. Peng has
affinity reporter ligands. This includes several instructive described in detail the use of cross-correlated 19F relaxation
simulations showing the relationships between the equilib- measurements for the study of ligand–receptor interactions
rium parameters – fraction bound, fractional reductions in and show how these data provide estimates of KD [108].
NMR responses etc., and the defining parameters of the
system – KD, KI, [P]0, [L1]0, and [L2]0 [102]. 8. Alternative measures of protein–ligand binding affinity

7. KD from ligand dissociation kinetics Although the dissociation constant KD is the preferred
quantitative measure of stability for bimolecular complexes
KD was defined earlier (Eq. (1)) in terms of the equilib- (because its meaning is clear), some other measures of affin-
rium concentrations of bound and free species. An equiva- ity are also used. It may not always be possible to define the
lent definition of KD is in terms of the equilibrium binding interaction in terms of a simple 1:1 complex. Some-
condition established by the balance of the ligand associa- times an experiment observable is related to KD, but in an
tion (kon) and dissociation (koff) rates indeterminate way, so that the response can only be used as
K D ¼ k on =k off : ð35Þ an indication of binding strength or as a ranking parame-
ter, but not as an absolute measure.
So instead of quantifying the solution speciation, one
might aim instead at measuring these rates. In very general
terms the first order ligand dissociation rate (units M1) is 8.1. Affinity index
a reflection of the strength of the intermolecular complex,
whereas the ligand association rate is a measure of how Rossi et al. have advocated the ‘affinity index’ as a mea-
quickly the ligand can arrive at the protein. An assumption sure of ligand macrocycle affinity [109]. Using the selective
is frequently made that the kon is diffusion limited and is spin–lattice relaxation rate R1 (1/T1(sel)) as the NMR obser-
thus independent of the system. A value of 1 · 109 M1 s1 vable, it can be shown that
is usually cited for kon. When this condition is satisfied koff  
1 1 1
is a useful proxy for the dissociation constant. There are ¼ þ ½L ; ð36Þ
DR KD R1ðboundÞ ½P0
many NMR protocols for measuring these rates [103].
Information on ligand exchange kinetics (koff and kon and therefore a plot of D1/T1(sel) versus the protein concen-
rates and hence KD) can be derived from complete line shape tration will be a straight line passing through the origin and
analysis of individual NMR peaks. Jardetzky et al. have with a slope
reported on a 1H NMR (500 MHz) study of the binding of  
L-tryptophan to the trp repressor of Escherichia Coli [52]. T K D R1ðboundÞ
½AL ¼ ; ð37Þ
This study is interesting because it used a full line shape anal- 1 þ K D ½L
L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242 239

which is defined as the ‘affinity index’. The dimensions of


T
½AL are M1 s1 and the T and L superscripts and sub-
scripts signify the temperature and ligand concentration
at which the measurement was made. The advantage of this
term is that it provides a measure of ligand-macromolecule
global affinity which is independent of the number of bind-
ing sites. A disadvantage of this parameter is its depen-
dence on the ligand concentration. So, although it can be
a useful way to rank order a series of ligands (or receptors)
within a single study where [L]0 can be controlled, it is
not a very efficient way to communicate knowledge of
receptor–ligand binding strength. The method has been
used to study the interaction of carbamazepine with
albumin [110], chloramphenicol and thiamphenicol with Fig. 14. IC50 of an experimental drug (known as H89) determined by 19F
albumin [111], and anandamide with multiple cannabinoid NMR detection of the CF3 labelled starting and enzymatically modified
receptors [112]. substrate. The test system comprised of a protein kinase and a 19F labelled
substrate peptide. This plot shows the unphosphorylated peptide concen-
tration, [pep], as a function of the inhibitor concentration [H89]. The best fit
8.2. Affinity ranking inhibition curve to the experimental data gives an IC50 of 0.72 ± 0.05 lM
[114]. Reproduced with permission.  2003 American Chemical Society.
NMR is now commonly used in the pharmaceutical
industry as a lead discovery tool in the drug discovery pro-
cess. Relatively small compound collections are screened of 1D 19F NMR data to quantify the amount of substrate
for receptor binding and the NMR response is used to sig- and the product of the enzyme reaction. The compound
nal binding. The usual outcome from such an endeavour is library is added to 384 well plates containing the active
a subset of ‘hits’ that have some affinity for the receptor. It enzyme and substrate and the reaction is quenched at a
is highly desirable to rank order these hits, but there may time when some fraction of the substrate is expected to
not be enough time or resource to apply the titration meth- be consumed. Automated processing of the 19F NMR spec-
ods. One crude but pragmatic way forward is to rank the tra is able to quickly identify the active compounds. The
ligands according to the magnitude of the NMR measured IC50 can be determined by collecting a full inhibition bind-
binding response. ing curve (Fig. 14), or it can be estimated approximately
A search for ligands of human adipocyte fatty acid bind- from a single point measurement because the values for
ing protein (FABP4) is a recent example. A 1H 1D T1q both plateaus of Fig. 14 are known from references and
relaxation filter experiment was used to identify ligands blanks.
from a collection of 531 compound which was initially Cryogenically cooled probes permit the application of
screened as cocktails of 5–10 compounds. The ligands were this kind of screening with femtomole levels of target
then classified as weak or strong binders according to the enzyme, a concentration similar to that required for tradi-
amplitude of the attenuation response in a second T1q tional high throughput screening methods [102].
relaxation filter experiment applied with a shorter spin lock
time [113]. 9. Applications of CP-MAS NMR

8.3. NMR as a functional biological screen No information about the equilibrium binding affinity is
available from solid crystalline protein–ligand complexes.
The requirements of the pharmaceuticals industry to However the techniques of solid state NMR can be usefully
invent more effective lead discovery programs has also applied to the hydrated gelatinous samples that are typical
led to the development of an NMR based functional of membrane bound protein preparations. Cross-polariza-
screen. A functional screen is one where tentative new tion magic angle spinning (CP-MAS) NMR goes some
drugs are tested against the fully functional target, e.g., way to resolving the technical difficulties of studying very
an enzyme that is actively turning over a substrate. The large proteins embedded in a lipid membrane. Add to this
method, termed 3-FABS (three fluorine atoms for bio- the simplification resulting from isotope labelled ligands,
chemical screening) by its inventors, uses NMR to analyse usually labelled at a single site with either 13C, 15N or
19
the progress of the enzyme reaction and is able to report F, and direct observation of the dynamics of receptor
the 50% mean inhibition concentration (IC50) of the active bound ligands becomes possible. Unlike the high resolution
ligand [114,115]. It is interesting that neither the receptor solution phase techniques which sample an exchange aver-
(the enzyme), nor the screened ligand are observed in this aged population, the CP-MAS experiment only sees the
experiment. The method requires the substrate of the ligand that is bound to the receptor. Thus the few reports
enzyme to be labelled with 19F. NMR analysis of the reac- of quantitation of KD by these methods are a special case
tion progress is then based on straightforward integration of ligand observed experiments.
240 L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

CP-MAS works because it distinguishes bound from


non-bound ligand. The 13C CP-MAS experiment generates
signals from 13C nuclei in rigid solids by transferring mag-
netization from abundant 1H nuclei. The experiment first
excites 1H magnetization and then transfers magnetization
from 1H to 13C during the contact time. The 1H-13C dipole
coupling is averaged to zero for rapidly reorienting small
molecules, so the 13C magnetization in the non-bound
ligand remains at equilibrium (weak), and is essentially
not detected. If at some point during the contact time
(spin-lock field applied) the ligand binds to the protein, it
will build up 13C magnetization at a predictable rate and
will produce a signal during the detection period. The
result is that the CP-MAS experiment can give a direct read
out of [PL] unperturbed by [L]. Measuring [PL] is the key
to knowing KD.
A study of the weak affinity binding of galactose to the Fig. 16. Simulations of cross polarization intensity profiles for a ligand
interacting with a membrane protein, calculated for different values of the
lactose transport protein LacS in native membranes is the
dissociation rate constant (koff) and the dissociation constant (KD). The
first reported estimation of KD from CP-MAS data [116]. virtual system was [L]0 = 6 mM, THC = 1.5 ms, T free 1qH ¼ 100 ms, and
Use of [1-13C]D-galactose allowed clear observation of the T bound
1qH ¼ 2 ms. The simulations represent [P] 0 = 2.4 mM (dashed line),
bound ligand against the background of natural abun- 2.0 mM (solid line) and 1.6 mM (dotted line) [117]. Reproduced with
dance carbon. Cross polarization NMR had not been permission.  2004 American Chemical Society.
thought to be amenable to quantitative interpretation since
the response of the observed nucleus is dependent on multi- The rate at which 13C magnetization builds up in the
ple and difficult to quantify interactions with nearby spins. ligand is a balance between the positive addition of magne-
However, a selectively bound substrate should experience a tization through the dipolar coupling and loss through
consistent environment throughout a titration, and the relaxation. These factors in turn depend on the binding
response can be analysed as a function of [L]0. This is constant and also the number of times that the ligand
shown in Fig. 15. The two data points at the highest [L]0 might exchange on and off the protein during the contact
were deemed to have been corrupted and were discarded time. Hence the response is a function of both KD and koff.
from the Scatchard plot. Note that the normalization step Simulated profiles of expected signal intensity are shown in
in this example assumes that saturation binding has Fig. 16. The experiment has been demonstrated with stud-
occurred at 5 mM ligand (certainly not the case) and is a ies of the binding of methyl [1-13C]-b-D-glucuronide to the
drastic simplification of the data analysis. GusB membrane transport protein from Escherichia coli
[117], and as a 19F NMR version for the anti-psychotic
drug trifluoroperazine binding to membrane embedded
gastric H+/K+ ATPase [118]. Under optimum conditions
the variable contact time CP-MAS experiment can be a sin-
gle point experiment. It is able to establish KD from a single
protein/ligand sample at just one protein:ligand ratio.

10. Conclusion

The use of NMR to determine quantitatively the associ-


ation constants of protein–ligand complexes is firmly estab-
lished. NMR might not always be the optimum means to
measure KD for a system, but NMR specialists obviously
enjoy pushing the boundaries and expanding the applica-
bility of magnetic resonance methods and the result is that
there are a lot of NMR experiments available for the task.
The tried and tested titration methods of sections 3 and 4
Fig. 15. Binding isotherm for [1-13C]-D-galactose to LacS membranes as are used very widely.
determined from the intensity of the substrate signal in the CP MAS 13C
The opportunity to observe cleanly the species of inter-
spectrum. The ratio of occupied binding sites was established by
normalizing the response to the maximum observed response (the response est with high sensitivity and often with no additional chem-
at 5 mM ligand). The inset is a Scatchard plot of the first four data points ical manipulation has resulted in the accumulation of huge
[116]. Reproduced with permission.  1999 American Chemical Society. experience with 19F NMR. Few would consider putting
L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242 241

fluorine into a ligand solely for the purpose of measuring [30] L.K. MacLachlan, D.G. Reid, R.C. Mitchell, C.J. Salter, S.J. Smith,
KD perhaps, but for systems where fluorine is already pres- J. Biol. Chem. 265 (1990) 9764.
[31] C. Dalvit, P. Floersheim, M. Zurini, A. Widmer, J. Biomol. NMR
ent in the ligand, the 19F NMR experiments might be con- 14 (1999) 23.
sidered as first choice methods for measuring KD. [32] C.J. Morton, D.J.R. Pugh, E.L.J. Brown, J.D. Kahmann, D.A.C.
The most recent developments with magnetization Renzoni, I.D. Campbell, Structure 4 (1996) 705.
transfer experiments, competition binding and CP-MAS [33] S.B. Shuker, P.J. Hajduk, R.P. Meadows, S.W. Fesik, Science 274
approaches, have resulted in novel, sensitive and specific (1996) 1531.
[34] P.J. Hajduk, G. Sheppard, D.G. Nettesheim, E.T. Olejniczak, S.B.
NMR methods to measure KD that have moved far from Shuker, R.P. Meadows, D.H. Steinman, G.M. Carrera, P.A.
the original linewidth and chemical shift perturbation Marcotte, J. Severin, K. Walter, H. Smith, E. Gubbins, R. Simmer,
approaches. They illustrate the breadth of interest in this T.F. Holzman, D.W. Morgan, S.K. Davidsen, J.B. Summers, S.W.
science and suggest that more interesting developments will Fesik, J. Am. Chem. Soc. 119 (1997) 5818.
continue to come. The most useful experiments will be [35] P.J. Hajduk, M. Bures, J. Praestgaard, S.W. Fesik, J. Med. Chem. 43
(2000) 3443.
those that establish [PL] directly and expeditiously. [36] H. Mao, P.J. Hajduk, R. Craig, R. Bell, T. Borre, S.W. Fesik, J. Am.
Chem. Soc. 123 (2001) 10429.
References [37] A.M. Petros, J. Dinges, D.J. Augeri, S.A. Baumeister, D.A.
Betebenner, M.G. Bures, S.W. Elmore, P.J. Hajduk, M.K. Joseph,
[1] D. Colquhoun, Trends Pharmacol. Sci. 27 (2006) 149. S.K. Landis, D.G. Nettesheim, S.H. Rosenberg, W. Shen, S.
[2] R.A. Dwek, Nuclear Magnetic Resonance (NMR) in Biochemistry, Thomas, X. Wang, I. Zanze, H. Zhang, S.W. Fesik, J. Med. Chem.
Clarendon Press, Oxford, 1973. 49 (2006) 656.
[3] G. Gemmecker, in: U. Holzgrabe, I. Wawer, B. Diehl (Eds.), NMR [38] H.-J. Boehm, M. Boehringer, D. Bur, H. Gmuender, W. Huber, W.
Spectroscopy in Drug Development and Analysis, Wiley-VCH, Klaus, D. Kostrewa, H. Kuehne, T. Luebbers, N. Meunier-Keller,
Weinheim, 1998, p. 135. F. Mueller, J. Med. Chem. 43 (2000) 2664.
[4] L. Fielding, Tetrahedron 56 (2000) 6151. [39] L. Parsons, N. Bonander, E. Eisenstein, M. Gilson, V. Kairys, J.
[5] K.A. Connors, Binding Constants, John Wiley & Sons, New York, Orban, Biochemistry 42 (2003) 80.
1987. [40] J.L. Weaver, J.H. Prestegard, Biochemistry 37 (1998) 116.
[6] V.M.S. Gil, N.C. Oliveira, J. Chem. Ed. 67 (1990) 473. [41] P.J. Hajduk, R. Mendoza, A.M. Petros, J.R. Huth, M. Bures, S.W.
[7] R.S. Macomber, J. Chem. Ed. 69 (1992) 375. Fesik, Y.C. Martin, J. Computer-Aided Mol. Design 17 (2003) 93.
[8] R.E. Barrans, D.A. Dougherty, Supramol. Chem. 4 (1994) 121. [42] B.J. Stockman, C. Dalvit, Prog. NMR Spectrosc. 41 (2002) 187.
[9] P. Gans, A. Sabatini, A. Vacca, Talanta 43 (1996) 1739. [43] E.R. Zartler, J. Yan, H. Mo, A.D. Kline, M.J. Shapiro, Curr. Topics
[10] G. Weber, S.R. Anderson, Biochemistry 4 (1965) 1942. Med. Chem. 3 (2003) 25.
[11] W.B. Person, J. Am. Chem. Soc. 87 (1965) 167. [44] B. Meyer, T. Peters, Angew. Chem. Int. Ed. 42 (2003) 864.
[12] D.A. Deranleau, J. Am. Chem. Soc. 91 (1969) 4044. [45] J.W. Peng, J. Moore, N. Abdul-Manan, Prog. NMR Spectrosc. 44
[13] D.A. Deranleau, J. Am. Chem. Soc. 91 (1969) 4050. (2004) 225.
[14] J. Granot, J. Magn. Reson. 55 (1983) 216. [46] J. Feeney, J.G. Batchelor, J.P. Albrand, G.C.K. Roberts, J. Magn.
[15] C.S. Wilcox, in: H.-J. Schneider, H. Dürr (Eds.), Frontiers in Reson. 33 (1979) 519.
Supramolecular Organic Chemistry and Photochemistry, VCH, [47] I.M. Klotz, Acc. Chem. Res. 7 (1974) 162.
Weinheim, 1991, p. 123. [48] S.J. Perkins, L.N. Johnson, D.C. Phillips, R.A. Dwek, Biochem. J.
[16] J.L. Sudmeier, J.L. Evelhoch, N.B.-H. Jonsson, J. Magn. Reson. 40 193 (1981) 553.
(1980) 377. [49] N.K. Sauter, M.D. Bednarski, B.A. Wurzburg, J.E. Hanson, G.M.
[17] A. Poveda, J. Jiménez-Barbero, Chem. Soc. Rev. 27 (1998) 133. Whitesides, J.J. Skehel, D.C. Wiley, Biochemistry 28 (1989) 8388.
[18] J.L. Asensio, F.J. Cañada, M. Bruix, A. Rodriguez-Romero, J. [50] T.J. Swift, R.E. Connick, J. Chem. Phys. 37 (1962) 307.
Jiménez-Barbero, Eur. J. Biochem. 230 (1995) 621. [51] K.A. Kronis, J.P. Carver, Biochemistry 24 (1985) 826.
[19] J.L. Asensio, F.J. Cañada, M. Bruix, C. González, N. Khiar, A. [52] T.H. Schmitt, Z. Zheng, O. Jardetzky, Biochemistry 34 (1995) 13183.
Rodrı́guez-Romero, J. Jiménez-Barbero, Glycobiology 8 (1998) 569. [53] F. Ni, Prog. NMR Spectrosc. 26 (1994) 517.
[20] J.F. Espinosa, J. L Asensio, J.L. Garcia, J. Laynez, M. Bruix, C. [54] J.J. Fischer, O. Jardetzky, J. Am. Chem. Soc. 87 (1965) 3237.
Wright, H.-C. Siebert, H.-J. Gabius, F.J. Cañada, J. Jiménez- [55] J.T. Gerig, J. Am. Chem. Soc. 90 (1968) 2681.
Barbero, Eur. J. Biochem. 267 (2000) 3965. [56] J. Miller, V. Witzemann, U. Quast, M.A. Raftery, Proc. Natl. Acad.
[21] J.L. Asensio, H.-C. Siebert, C.-W. von der Lieth, J. Laynez, M. Sci. USA 76 (1979) 3580.
Bruix, U.M. Soedjanaamadja, J.J. Beintema, F.J. Cañada, H.-J. [57] L. Verdier, J. Gharbi-Benarous, G. Bertho, N. Evrard-Todeschi, P.
Gabius, J. Jiménez-Barbero, Proteins 40 (2000) 218. Mauvais, J.-P. Girault, J. Chem. Soc. Perkin Trans. 2 (2000) 2363.
[22] N. Aboitiz, M. Vila-Perelló, P. Groves, J.L. Asensio, D. Andreu, [58] K.A. Kronis, J.P. Carver, Biochemistry 21 (1982) 3050.
F.J. Cañada, J. Jiménez-Barbero, Chem. Bio. Chem. 5 (2004) 1245. [59] L. Fielding, D. Fletcher, S. Rutherford, J. Kaur, J. Mestres, Org.
[23] A. De Marco, S.M. Hochschwender, R.A. Laursen, M. Llinás, J. Biomol. Chem. 1 (2003) 4235.
Biol. Chem. 257 (1982) 12716. [60] G. Valensin, T. Kushnir, G. Navon, J. Magn. Reson. 46 (1982) 23.
[24] A. De Marco, A.M. Petros, R.A. Laursen, M. Llinás, Eur. Biophys. [61] R.W. Behling, T. Yamane, G. Navon, M.J. Sammon, L.W. Jelinski,
J. 14 (1987) 359. Biophys. J. 53 (1988) 947.
[25] A.M. Petros, V. Ramesh, M. Llinás, Biochemistry 28 (1989) 1368. [62] Y. Fraenkel, G. Navon, A. Aronheim, J.M. Gershoni, Biochemistry
[26] T. Thewes, K. Constantine, I.-J.L. Byeon, M. Llinás, J. Biol. Chem. 29 (1990) 2617.
265 (1990) 3906. [63] W.S. Price, Aust. J. Chem. 56 (2003) 855.
[27] I.-J.L. Byeon, R.F. Kelly, M.G. Mulkerrin, S.S.A. An, M. Llinás, [64] L.H. Lucas, C.K. Larive, Concepts Magn. Reson. 20A (2004) 24.
Biochemistry 34 (1995) 2739. [65] T.S. Derrick, E.F. McCord, C.K. Larive, J. Magn. Reson. 155
[28] D.N. Marti, C.-K. Hu, S.S.A. An, P. von Haller, J. Schaller, M. (2002) 217.
Llinás, Biochemistry 36 (1997) 11591. [66] A.J. Lennon, N.R. Scott, B.E. Chapman, P.W. Kuchel, Biophys. J.
[29] R. Wintjens, J.-M. Wieruszeski, H. Drobecq, P. Rousselot-Pailley, 67 (1994) 2096.
L. Buées, G. Lippens, I. Landrieu, J. Biol. Chem. 276 (2001) 25150. [67] R.J. Labotka, C.M. Schwab, Anal. Biochem. 191 (1990) 376.
242 L. Fielding / Progress in Nuclear Magnetic Resonance Spectroscopy 51 (2007) 219–242

[68] M.L. Tillett, M.A. Horsfield, L.-Y. Lian, T.J. Norwood, J. Biomol. [95] B.G. Jenkins, Life Sci. 48 (1991) 1227.
NMR 13 (1999) 223. [96] J.T. Gerig, Prog. NMR Spectrosc. 26 (1994) 293.
[69] J. Fejzo, C.A. Lepre, J.W. Peng, G.W. Bemis, M.A. Murcko, J.M. [97] M.I. Chávez, C. Andreu, P. Vidal, N. Aboitiz, F. Freire, P. Groves,
Moore, Chem. Biol. 6 (1999) 755. J.L. Asensio, G. Asensio, M. Muraki, F.J. Cañada, J. Jiménez-
[70] A.P. Campbell, B.D. Sykes, J. Magn. Reson. 93 (1991) 77. Barbero, Chem. Eur. J. 11 (2005) 7060.
[71] A. Chen, M.J. Shapiro, J. Am. Chem. Soc. 122 (2000) 414. [98] K.L. Gammon, S.H. Smallcombe, J.H. Richards, J. Am. Chem. Soc.
[72] M. Mayer, B. Meyer, Angew. Chem. Int. Ed. 38 (1999) 1784. 94 (1972) 4573.
[73] V.V. Krishnan, Curr. Anal. Chem. 1 (2005) 307. [99] J. Scheuring, J. Lee, M. Cushman, H. Patel, D.A. Patrick, A.
[74] M. Mayer, B. Meyer, J. Am. Chem. Soc. 123 (2001) 6108. Bacher, Biochemistry 33 (1994) 7634.
[75] R. Meinecke, B. Meyer, J. Med. Chem. 44 (2001) 3059. [100] B.W. Dubois, A.S. Evers, Biochemistry 31 (1992) 7069.
[76] A.T. Neffe, M. Bilang, B. Meyer, Org. Biomol. Chem. 4 (2006) 3259. [101] C. Dalvit, P.E. Fagerness, D.T.A. Hadden, R.W. Sarver, B.J.
[77] A. Blume, A.J. Benie, F. Stolz, R.R. Schmidt, W. Reutter, S. Stockman, J. Am. Chem. Soc. 125 (2003) 7696.
Hinderlich, T. Peters, J. Biol. Chem. 31 (2004) 55715. [102] C. Dalvit, N. Mongelli, G. Papeo, P. Giordano, M. Veronesi, D.
[78] A.J. Benie, A. Blume, R.R. Schmidt, W. Reutter, S. Hinderlich, T. Moskau, R. Kümmerle, J. Am. Chem. Soc. 127 (2005) 13380.
Peters, J. Biol. Chem. 31 (2004) 55722. [103] O. Monasterio, Methods 24 (2001) 97.
[79] C. Dalvit, G. Fogliatto, A. Stewart, M. Veronesi, B. Stockman, J. [104] M.V. Krishna Sastry, M.J. Swamy, A. Surolia, J. Biol. Chem. 263
Biomol. NMR 21 (2001) 349. (1988) 14862.
[80] T. Peters, All About Albumin: Biochemistry, Genetics and Medical [105] M. Hensmann, G.W. Booker, G. Panayotou, J. Boyd, J. Linacre, M.
Applications, Academic Press, San Diego, 1996. Waterfield, I.D. Campbell, Protein Sci. 3 (1994) 1020.
[81] M. Liu, J.K. Nicholson, J.C. Lindon, Anal. Commun. 34 (1997) 225. [106] U.L. Günther, B. Schaffhausen, J. Biomol. NMR 22 (2002) 201.
[82] M. Tanaka, Y. Asahi, S. Masuda, T. Ota, Chem. Pharm. Bull. 37 [107] J. Scheuring, M. Fischer, M. Cushman, J. Lee, A. Bacher, H.
(1989) 3177. Oschkinat, Biochemistry 35 (1996) 9637.
[83] M. Tanaka, Y. Asahai, S. Masuda, T. Ota, Chem. Pharm. Bull. 39 [108] J.W. Peng, J. Magn. Reson. 153 (2001) 32.
(1991) 2771. [109] C. Rossi, A. Donati, C. Bonechi, G. Corbini, R. Rappuoli, E.
[84] R.-S. Luo, M.-L. Liu, X.-A. Mao, Appl. Spectrosc. 53 (1999) 776. Dreassi, P. Corti, Chem. Phys. Lett. 264 (1997) 205.
[85] C.-G. Li, M.-L. Liu, C.-H. Ye, Appl. Magn. Reson. 19 (2000) 179. [110] C. Rossi, C. Bonechi, S. Martini, M. Ricci, G. Corbini, P. Corti, A.
[86] L. Fielding, S. Rutherford, D. Fletcher, Magn. Reson. Chem. 43 Donati, Magn. Reson. Chem. 39 (2001) 457.
(2005) 463. [111] S. Martini, C. Bonechi, A. Magnani, N. Marchettini, P. Corti, G.
[87] Y.F. Cui, G.Y. Bai, C.G. Li, C.H. Ye, M.L. Liu, J. Pharm. Biomed. Corbini, C. Rossi, Magn. Reson. Chem. 41 (2003) 489.
Anal. 34 (2004) 247. [112] C. Bonechi, S. Martini, V. Brizzi, P. Massarelli, G. Bruni, C. Rossi,
[88] R-S. Luo, M.-L. Liu, X.-A. Mao, Spectrochim. Acta A 55 (1999) 1897. Eur. J. Med. Chem. 41 (2006) 1117.
[89] W.S. Price, F. Elwinger, C. Vigouroux, P. Stilbs, Magn. Reson. [113] M.J.P. van Dongen, J. Uppenberg, S. Svensson, T. Lundbäck, T.
Chem. 40 (2002) 391. Åkerud, M. Wikström, J. Schultz, J. Am. Chem. Soc. 124 (2002)
[90] J.E. Hanson, N.K. Sauter, J.J. Skehel, D.C. Wiley, Virology 189 11874.
(1992) 525. [114] C. Dalvit, E. Ardini, M. Flocco, G.P. Fogliatto, N. Mongelli, M.
[91] C. Dalvit, M. Fasolini, M. Flocco, S. Knapp, P. Pevarello, M. Veronesi, J. Am. Chem. Soc. 125 (2003) 14620.
Veronesi, J. Med. Chem. 45 (2002) 2610. [115] C. Dalvit, E. Ardini, G.P. Fogliatto, N. Mongelli, M. Veronesi,
[92] C. Dalvit, M. Flocco, S. Knapp, M. Mostardini, R. Perego, B.J. Drug Discov. Today 9 (2004) 595.
Stockman, M. Veronesi, M. Varasi, J. Am. Chem. Soc. 124 (2002) [116] P.J.R. Spooner, L.M. Veenhoff, A. Watts, B. Poolman, Biochemistry
7702. 38 (1999) 9634.
[93] Y.-S. Wang, D. Liu, D.F. Wyss, Magn. Reson. Chem. 42 (2004) 485. [117] S.G. Patching, A.R. Brough, R.B. Herbert, J.A. Rajakarier, P.J.F.
[94] W. Jahnke, P. Floersheim, C. Ostermeier, X. Zhang, R. Hemmig, K. Henderson, D.A. Middleton, J. Am. Chem. Soc. 126 (2004) 3072.
Hurth, D.P. Uzunov, Angew. Chem. Int. Ed. 41 (2002) 3420. [118] M.P. Boland, D.A. Middleton, Magn. Reson. Chem. 42 (2004) 204.

Вам также может понравиться