You are on page 1of 9

NMR Spectroscopy, 31P

David G Gorenstein and Bruce A Luxon, University of Texas Medical Branch, Galveston, TX, USA
ã 2017 Elsevier Ltd. All rights reserved.

Symbols Du change in the s-bond angle


a2 percentage s character DxX electronegativity difference in the P–X bond
J coupling constant s shielding constant
S overlap integral siso isotropic shielding constant
t1 delay time sk parallel component of shielding constant
t2 observe time in 2D NMR s? perpendicular component of shielding
T1 spin–lattice relaxation time constant
T2 time constant for the FID s11, s22, s33 components of shielding tensor
Dd chemical shift difference f dihedral angle

Introduction Practical Considerations

Although 31P NMR spectra were reported as early as 1951, it Today’s commercial NMR spectrometers cover the 31P fre-
was the availability of commercial multinuclear NMR spec- quency range from 24 to 323 MHz (on a 800 MHz 1H NMR
trometers in about 1955 that led to the application of 31P spectrometer). Generally, for small, phosphorus-containing
NMR as an important analytical tool for structure elucidation. compounds, high signal-to-noise and resolution requirements
Early spectrometers generally required neat samples in large dictate use of as high a magnetic field strength as possible since
nonrotating tubes (8–12 mm OD). By the mid-1960s NMR both sensitivity and chemical-shift dispersion increase at
became more sensitive and the availability of higher field higher operating frequency (and field). However, consider-
electromagnets and signal averaging spurred rapid growth in ation must be given to field-dependent relaxation mechanisms
the number of reported 31P spectra and the publication of the such as chemical shift anisotropy which can lead to substantial
first monograph devoted to this field. line broadening of the 31P NMR signal at high fields. Indeed,
With the introduction of Fourier-transform (FT) and high- especially for larger biomolecules, sensitivity is often poorer at
field superconducting-magnet NMR spectrometers in about very high fields because of considerably increased line widths.
1970, 31P NMR spectroscopy expanded beyond the study of The latest probe designs provide a remarkable improvement in
small organic, organometallic and inorganic compounds to signal-to-noise. Indirect detection 2D NMR experiments have
biological phosphorus-containing compounds as well. The further improved sensitivity.
latest multinuclear FT NMR spectrometers have reduced if not Typical acquisition times for the 1D 31P free induction
eliminated the serious limitation to the widespread utilization decay (FID) following a 90 radiofrequency pulse are 1–8 s
of phosphorus NMR, which is the low sensitivity of the phos- depending on the required resolution (dictated by the line
phorus nucleus (6.6% at constant field compared to 1H NMR; width of the signal, 1/p T2 , where T2 is the time constant
magnetogyric ratio g, 10.839107 rad T1 s1; NMR frequency for the FID). Waiting longer than 2T2 will generally not
at 4.7 T, 80.96 MHz). Today, routinely, millimolar (or lower) improve the signal-to-noise ratio (S/N). Additional consider-
concentrations of phosphorus nuclei in as little as 0.3 mL of ation for optimization of the S/N must be given to the time it
solution are conveniently monitored. The 31P nucleus has takes for the 31P spins to return to thermal equilibrium after
other convenient NMR properties making it suitable for FT a 90 radiofrequency pulse, which is roughly 3 times the
NMR: spin 12 (which avoids problems associated with quadru- spin–lattice relaxation time (T1). To obtain 99.3% relaxation,
polar nuclei), 100% natural abundance, moderate relaxation a delay of 5 times T1 is necessary. If T1T2 , as would be true
times (providing relatively rapid signal averaging and sharp for small phosphorus-containing molecules where magnetic
lines), and a wide range of chemical shifts (>2000 ppm). field inhomogeneity and paramagnetic impurities do not lead
In this article the interpretation of various 31P NMR spec- to any additional line broadening, then a waiting period
troscopic parameters, particularly chemical shifts and coupling between pulses of 3T2 provides a good compromise between
constants, will be described. A major emphasis will be placed adequate resolution and signal sensitivity. If T1>T2 , as is often
on developments in 31P NMR methods which have consider- the case in larger biomolecular systems, then waiting only 3T2
ably expanded the utility of this important spectroscopic probe does not allow the magnetization to return to equilibrium and
in organic and biological structure determination. an additional delay must generally be introduced so that the
total time between pulses is 3T1. This wait can be substan-
tially shortened if the Ernst relationship is used to set the pulse
flip angles to <90 . At low field, 60–70 pulses, 4 to 8 k data
This article is reproduced from the previous edition, Copyright points and 2.0–5.2 s recycle times are generally used. The spec-
1999, Elsevier Ltd. tra are generally broadband 1H decoupled.

294 Encyclopedia of Spectroscopy and Spectrometry, Third Edition http://dx.doi.org/10.1016/B978-0-12-803224-4.00242-9


NMR Spectroscopy, 31P 295

The 31P spectra are generally referenced to an external sam- 1


siso ¼ ðs11 þ s22 þ s33 Þ [2]
ple of 85% H3PO4 or trimethylphosphate which is 3.46 ppm 3
downfield of 85% H3PO4. Note that throughout this review and the anisotropy Ds is given by
the IUPAC convention is followed so that positive values are to
high frequency (low field). One should cautiously interpret 1
Ds ¼ s11  ðs22 þ s33 Þ [3]
reported 31P chemical shifts because the early literature (pre- 2
1970s) and even many later papers use the opposite sign or, for axial symmetry,
convention.
Ds ¼ sk  s? [4]

Quantification of Peak Heights


Theoretical 31P Chemical Shift Calculations and Empirical
Historically, the intensity of a resonance has been measured in Observations
several ways: (1) peak heights and areas obtained from the
standard software supplied by the spectrometer manufacturer, Three factors appear to dominate 31P chemical shift differences
(2) peak heights measured by hand, (3) peaks cut and weighed Dd, as shown by
from the plotted spectrum and (4) peaks fitted to a Lorentzian Dd ¼ CDwX þ kDnp þ ADy [5]
line shape. For flat baselines, intensity measurements are gen-
erally straightforward. However, in the event of curved base- where DwX is the difference in electronegativity in the P–X
lines the measurements are somewhat uncertain and careful bond, Dnp is the change in the p-electron overlap, Dy is the
baseline correction is necessary. change in the s-bond angle, and C, k, and A are constants.
It is often necessary that experiments be carried out As suggested by eqn [5], electronegativity effects, bond
without allowing time for full recovery of longitudinal magne- angle changes, and p-electron overlap differences can all poten-
tization between transients because of the limited availability tially contribute to 31P shifts in a number of classes of phos-
of spectrometer time or of the limited lifetime of the sample. phorus compounds. While these semiempirical isotropic
Because of variations in T1 between different phosphates chemical-shift calculations are quite useful in providing a
and variation in the heteronuclear NOE to nearby protons, chemical and physical understanding for the factors affecting
31
care should be made in interpretation of peak area and inten- P chemical shifts, they represent severe theoretical approxi-
sities. Addition of a recycle delay of at least 5T1 between mations. More exact ab initio chemical-shift calculations of the
pulses and gated decoupling only during the acquisition time shielding tensor are very difficult although a number of calcu-
to eliminate the 1H–31P NOE largely eliminates quantification lations have been reported on phosphorus compounds.
problems. Whereas the semi-empirical theoretical calculations have
largely supported the importance of electronegativity, bond
angle, and p-electron overlap on 31P chemical shifts, the equa-
tions relating 31P shift changes to structural and substituent
31 changes unfortunately are not generally applicable. Also,
P Chemical Shifts
because 31P shifts are influenced by at least these three factors,
Introduction and Basic Principles empirical and semiempirical correlations can only be applied
The interaction of the electron cloud surrounding the phos- to classes of compounds that are similar in structure. It should
phorus nucleus with an external applied magnetic field B0 gives also be emphasized again that structural perturbations will
rise to a local magnetic field. This induced field shields the affect 31P chemical shift tensors. Often variations in one of the
nucleus, with the shielding proportional to the field B0 so that tensor components will be compensated for by an equally large
the effective field, Beff, felt by the nucleus is given by variation in another tensor component with only a small net
effect on the isotropic chemical shift. Interpretation of varia-
Beff ¼ B0 ð1  sÞ [1] tions of isotropic 31P chemical shifts should therefore be
where s is the shielding constant. Because the charge distribu- approached with great caution.
tion in a phosphorus molecule will generally be far from Within these limitations, a number of semiempirical and
spherically symmetrical, the 31P chemical shift (or shielding empirical observations and correlations, however, have been
constant) varies as a function of the orientation of the mole- established and have proved useful in predicting 31P chemical-
cule relative to the external magnetic field. This gives rise to a shift trends. Indeed, unfortunately, no single factor can readily
chemical-shift anisotropy that can be defined by three princi- rationalize the observed range of 31P chemical shifts (Figure 1).
pal components, s11, s22 and s33, of the shielding tensor. For
molecules that are axially symmetrical, with s11 along the Bond Angle Effects
principal axis of symmetry, s11¼sk (parallel component), Changes in the s-bond angles appear to make a contribution
and s22¼s33¼s? (perpendicular component). These aniso- (|A|, eqn [5]) to the 31P chemical shifts of phosphoryl com-
tropic chemical shifts are observed in solid samples and liquid pounds, although electronegativity effects apparently pre-
crystals, whereas for small molecules in solution, rapid tum- dominate. Empirical correlations between 31P chemical
bling averages the shift. The average, isotropic chemical shield- shifts and X–P–X bond angles can be found, although success
ing siso (which would be comparable to the solution chemical here depends on the fact that these correlations deal with
shift) is given by the trace of the shielding tensor or only a limited structural variation: in the case of phosphate
296 NMR Spectroscopy, 31P

Figure 1 Typical 31P chemical shift ranges for phosphorus bonded to various substituents in different oxidation states. (P– indicates the P4 molecule.)

esters, it is the number and chemical type of R groups (correlation spectroscopy via long range coupling) experiment
attached to a tetrahedron of oxygen atoms surrounding the and indirect detection (1H detection) HETCOR experiments can
phosphorus nucleus. For a wide variety of different alkyl be used to assign multiple 31P signals in complex spectra such as
phosphates (mono-, di-, and triesters, cyclic and acyclic neu- those of oligonucleotide duplexes. Additional 2D heteronuclear
tral, monoanionic, and dianionic esters), at bond angles J cross-polarization TOCSY (TOCSY¼total correlation spectros-
<108 a decrease in the smallest O–P–O bond angle in the copy), 2D heteronuclear TOCSY-NOESY (NOESY¼nuclear
molecule generally results in a deshielding (downfield shift) Overhauser effect spectroscopy), and even a 3D heteronuclear
of the phosphorus nucleus. TOCSY-NOESY experiment can be used if additional spectral
dispersion, by adding a third frequency dimension, is desirable.
Torsional Angle Effects on 31P Chemical Shifts This may prove to be extremely valuable for ribo-
Semi-empirical molecular orbital calculations and ab initio oligonucleotides where very little 1H spectral dispersion in the
gauge-invariant-type molecular orbital, chemical-shift calcula- sugar proton chemical shifts is unfortunately observed.
tions suggested that 31P chemical shifts are also dependent on Generally these 2D experiments correlate 31P signals with
P–O ester torsional angles and this has been shown to be of coupled 1H NMR signals. Assuming the 1H NMR spectra have
great value in analysis of DNA structure (see below). The two been assigned, these methods allow for direct assignment of
nucleic acid P–O ester torsional angles, z (50 -O–P) and a the 31P signals. The HETCOR measurements, however, suffer
(30 -O–P), are defined by the (50 -O–P–O-30 ) backbone dihedral from poor sensitivity as well as poor resolution in both the
1
angles. These chemical-shift calculations and later empirical H and 31P dimensions, especially for larger biomolecular
observations indicated that a phosphate diester in a BI confor- structures. The poor sensitivity is largely due to the fact that
mation (both ester bonds gauche() or 60 ) should have a the 1H–31P scalar coupling constants are generally about the
31
P chemical shift 1.6 ppm upfield from a phosphate diester in same size or smaller (except for organophosphorus molecules
the BII conformations (a¼gauche(); z¼trans or 180 ). with directly bonded hydrogens) than the 1H–1H coupling
constants. Sensitivity is substantially improved by using a het-
eronuclear version of the ‘constant time’ coherence transfer
31
P Signal Assignments technique, referred to as COLOC and originally proposed for
13
C–1H correlations.
If the proton spectra of the molecule has been previously An example of a 2D HETCOR spectrum of the self-
assigned, then 2D 31P–1H heteronuclear correlation NMR complementary 14-base-pair oligonucleotide duplex, d
spectroscopy can generally provide the most convenient (TGTGAGCGCTCACA)2, is shown in Figure 2. The cross-
method for assigning 31P chemical shifts in complex spectra. peaks represent scalar couplings between 31P nuclei of the
Whilst application of these experiments to DNA is clear, the 2D backbone and the H30 and H40 deoxyribose protons. Assuming
methods will of course equally apply to organophosphorus that the chemical shifts of these protons have been assigned
compounds as well. (by 1H–1H NOESY and COSY spectra) the 31P signals may be
Conventional 2D 31P–1H heteronuclear shift correlation readily assigned (COSY¼homonuclear chemical shift correla-
(HETCOR) NMR spectroscopy, the 2D long-range COLOC tion spectroscopy).
NMR Spectroscopy, 31P 297

Figure 2 Pure absorption phase 31P–1H heteronuclear correlation spectrum of tetradecamer duplex d(TGTGAGCGCTCACA)2 at 200 MHz (1H).
31
P chemical shifts are reported relative to trimethyl phosphate which is 3.456 ppm downfield from 85% phosphoric acid. Reproduced
with permission.

Coupling Constants
1
JPC. Calculations and empirical observations on trivalent
phosphorus compounds are not successful however, and sug-
Directly Bonded Phosphorus Coupling Constants 1JPX gest that the Fermi-contact contribution only dominates tet-
One bond P–X coupling constants (JPX) have generally been ravalent phosphorus compounds.
rationalized in terms of a dominant Fermi-contact term One-bond P–H coupling constants appear always to be
positive and vary from about þ120 to þ1180 Hz. Other het-
Aa2P a2X eroatom one-bond P–X coupling constants vary over a similar
JPX ¼ þB [6]
1 þ S2PX wide range and can be either positive or negative. The expected
range of values is given in Table 1.
where A and B are constants, a2P and a2X are percentage s
character on phosphorus and atom X, respectively, and SPX
is the overlap integral for the P–X bond. Because the Fermi- Two Bond Coupling Constants: 2JPX
contact spin–spin coupling mechanism involves the electron
Two-bond 2JPX coupling constants may be either positive or
density at the nucleus (hence the s-orbital electron density),
negative and are generally smaller than one-bond coupling
an increase in the s character of the P–X bond is generally
constants (Table 2). The 2JPCH and 2JPCF constants are stereo-
associated with an increase in the coupling constant. The
specific and a Karplus-like dihedral dependence to the two-
percentages character is determined by the hybridization of
bond coupling constant (H or F)–C–P–X (X¼lone pair or
atoms P and X, and as expected sp3-hybridized atoms often
heteroatom) has been found. Thus in the cis- and trans-phos-
have 1JPX larger than p3 hybridized atoms. Thus 1JPH for
phorinanes, the 2JPC constants are 0.0 and 5.1 Hz in the cis- and
phosphonium cations of structure PHnRþ 3
4n with sp hybrid-
trans-isomers, respectively.
ization are 500 Hz, whereas JPH for phosphines PHnR3n
1

with phosphorus hybridization of approximately p3 are smal-


ler, 200 Hz. Furthermore, as the electronegativity of atom X Three-Bond Coupling Constants, 3JPX
increases, the percentages character of the P–X bond
increases, and the coupling constant becomes more positive. Three-bond coupling constant, 3JPX, through intervening C, N,
In many cases, however, these simple concepts fail to ratio- O, or other heteroatoms are generally <20 Hz (Table 3). The
nalize experimental one-bond P–X coupling constants dihedral-angle dependence of vicinal 3JPOCH coupling 3JPCCH
(Table 1) because other spin–spin coupling mechanisms and 3JPCCC has been demonstrated. The curves may be fitted to
can also contribute significantly to the coupling constant. the general Karplus equation
For tetravalent phosphorus, a very good correlation is
JðfÞ ¼ A cos 2 f þ B cos f þ C [7]
found between 1JPC and the phosphorus 3s–carbon 2s
bond orders, the percentage s in the P–C bonding orbital in where f is the dihedral angle and A, B and C are constants for
going from alkyl to alkenyl to alkynyl (sp3!sp2!sp), and the particular molecular framework. Caution is recommended
298 NMR Spectroscopy, 31P

Table 1 One-bond phosphorus spin–spin coupling constants 1JPX Table 2 Two-bond phosphorus spin–spin coupling constants 2JPX
1
Structural class (or structure) J (HZ)a Structural class (or structure) 2
J (Hz)a

P(II) P(III)
PH2 139
180–225 0–18

0–45
P(CH3)3 þ2.7

820–1450
40–149

100–400
P(CF3)3 85.5

P(IV)
13–28
490–600

12–20
50–305

P(C2H5)3 þ14.1
Pþ(CH3)4 þ56
460–1030
10–12

1000–1400
(M¼O, S) 70–90
(X¼S, C)
490–650
P(IV)

P(V)
7–30

700–1000
P(O)(CH3)3 12.8, 13.4

530–1100
12–18

PF5 938 0–40

P(VI) Pþ(C2H5)4 4.3

PF
6 706 6
a
For structural classes, only the absolute value for J is given.
P(V)

10–18

when attempting to apply these Karplus equations and curves 124–193


to classes of phosphorus compounds that have not been used
in establishing these relationships because separate correla-
tions and values for the constants A, B and C in eqn [7] P(VI)
probably exist for each structural class. In all cases, a minimum
in these Karplus curves is found at 90 . 130–160

Applications to Nucleic Acid Structure


a
For structural classes, only the absolute value for J is given.
The Karplus-like relationship between HCOP and CCOP dihe-
dral angles and 3JHP and 3JCP three-bond coupling constants,
respectively, has been used to determine the conformation
NMR Spectroscopy, 31P 299

Table 3 Three-bond phosphorus spin–spin coupling constants 3JPX The JH30 P coupling constants in larger oligonucleotides
3
cannot generally be determined from the coupled 1D 31P or
Structural class (or structure) J (Hz)a 1
H spectra because of spectral overlap. 2D J-resolved long-
P(III) range correlation pulse sequences can be used to overcome
this limitation. The Bax–Freeman selective 2D J experiment
with a DANTE (delays alternating with nutations for tailored
0–15 excitation) sequence for a selective 180 pulse on the coupled
protons can be readily implemented on most spectrometers.
P(OCH3)3 10.8–11.8 This is particularly useful for measuring phosphorus–H30 cou-
pling constants in duplex fragments, which can vary from 1.5
to 8 Hz in duplexes as large as tetradecamers. There is a strong
10–16
correlation (R¼0.92) between torsional angles
C40 –C30 –O30 –P (e) and C30 –O30 –P–O50 (z) in the crystal struc-
3–14 tures of various duplexes. Thus both torsional angles e and z
can often be calculated from the measured P–H30 coupling
P[N(CH3)2]3 8.8–9.0 constant.
Coupling constants of both 50 protons are analysed in order
to determine conformations about the C50 –O bond. Unfortu-
P(IV)
nately, these b torsional angles have in practice been generally
unobtainable even in moderate-length duplexes. Selective 2D
7–11 J-resolved spectra generally fail for H40 , H50 or H500 coupling to
31
P because the spectral dispersion between these protons is so
15–22 limited. However, with either 13C labelling or even natural
abundance 13C methods, it is possible to measure not only
the 1H–31P but also the 13C–31P coupling constants. Analysis
16–20
of the 2D multiplet pattern, especially the ‘E. COSY’ pattern of
the 1H–13C HSQC spectrum, has allowed extraction of many
0–13 carbon (C30 , C40 , C50 ) and proton (H30 , H40 , H50 , H500 ) cou-
pling constants to phosphorus. The larger line widths of longer
P(O)(OCH3)3 10.2–11.4 duplexes limit measurement of the small coupling constants.
As shown in Figure 3, the Karplus relationship provides for
four different torsional angle solutions for each value of the
14–25 same coupling constant. Although all four values are shown in
(M¼O, S)
Figure 3, the limb which includes e values between 360 and
270 is sterically inaccessible in nucleic acids. As shown in
Figure 3, nearly all of the phosphates for normal Watson–Crick
duplexes fall along only a single limb of the Karplus curve.
P(V)
Thus, for ‘normal’ B-DNA geometry, there is an excellent cor-
relation between the phosphate resonances and the observed
20–27 torsional angle, while phosphates that are greatly distorted in
their geometry must be more carefully analysed.
12–17 It is clear from Figure 3 that 31P chemical shifts and cou-
pling constants provide probes of the conformation of the
phosphate ester backbone in nucleic acids and nucleic acid
a
complexes. It is important to remember that 31P chemical
For structural classes, only the absolute value for J is given.
shifts are dependent on factors other than torsional angles
alone. As noted above, 31P chemical shifts are very sensitive
to bond angle distortions as well. It is quite reasonable to
assume that backbone structural distortions as observed in
about the ribose-phosphate backbone of nucleic acids in solu- unusual nucleic acid structures also introduce some bond
tion. Torsional angles about both the C30 –O30 and C50 –O50 angle distortion as well. Widening of the ester O–P–O bond
bonds in 30 ,50 -phosphodiester linkages have been determined angle indeed is expected to produce an upfield shift, while
from the coupled 1H and 31P NMR spectra. narrowing of this bond angle causes a downfield shift, and it
Within the limitations just described for the general appli- is possible that this bond angle effect could account for the
cation of the Karplus relationship, the best Karplus relation- anomalous shifts. Indeed, very large 31P chemical shift varia-
ship for the nucleotide H30 –P coupling constants appears to be tions (3–7 ppm) are observed in transfer RNA and hammer-
head RNA phosphates, and are probably due to bond angle
J ¼ 15:3 cos 2 ðyÞ  6:1 cos ðyÞ þ 1:6
distortions in these tightly folded structures.
From the H30 –C30 –O–P torsional angle y, the C40 –C30 –O–P Generally the main-chain torsional angles of the individual
torsional angle e (¼y120 ) may be calculated. phosphodiester groups along the oligonucleotide double helix
300 NMR Spectroscopy, 31P

Figure 3 Plot of 31P chemical shifts for duplex oligonucleotide sequences (○) and an actinomycin D bound d(CGCG)2 tetramer complex (□) with
measured JH30 P coupling constants (•, phosphates in a tandem GA mismatch decamer duplex which shows unusual, slowly exchanging signals).
Also shown are the theoretical e and z torsion angles (solid curve) as a function of the coupling constant derived from the Karplus relationship
(e) and the relationship z¼3171.23e. 31P chemical shifts are reported relative to trimethyl phosphate. Reproduced with permission.

are responsible for sequence-specific variations in the 31P Table 4 Chemical shifts and pH titration data for representative
chemical shifts. In duplex B-DNA, the gauche(), gauche() model compounds
(g, g; z, a) (or BI) conformation about the P–O ester bonds
in the sugar phosphate backbone is energetically favoured, and Compound Chemical shift Titratableb pKa
this conformation is associated with a more shielded 31P reso- (ppm)a
nance. In both duplex and single stranded DNA the trans, Phosphomonoesters
gauche() (t, g; z, a) (or BII) conformation is also signifi- Phosphoserine 4.6 þ 5.8
cantly populated. Phosphothreonine 4.0 þ 5.9
The 31P chemical shift difference between the BI and BII Pyridoxal phosphate 3.7 þ 6.2
phosphate ester conformational stages is estimated to be Pyridoxamine 3.7 þ 5.7
1.5–1.6 ppm. As the result of this sensitivity to the backbone phosphate
conformational state, 31P chemical shifts of duplex oligonucle- Flavin mononucleotide 4.7 þ 6.0
otides have been shown to be dependent upon both the Phosphodiesters
RNA, DNA, 0 to 1.5 
sequence and the position of the phosphate residue.
phospholipids
The possible basis for the correlation between local helical
Diphosphodiesters
structural variations and 31P chemical shifts can be analysed in Flavin adenine 10.8 to 11.3 
terms of deoxyribose phosphate backbone changes involved in dinucleotide
local helical sequence-specific structural variations. As the helix Phosphotriesters
winds or unwinds in response to local helical distortions, the Dialkyl phosphoserine 0 to 3.0 
length of the deoxyribose phosphate backbone must change to Phosphoramidates
reflect the stretching and contracting of the deoxyribose phos- N3-Phosphohistidine 4.5 
phate backbone between the two stacked base pairs. To a N1-Phosphohistidine 5.5 
significant extent, these changes in the overall length of the Phosphoarginine 3.0 þ 4.3
Phosphocreatine 2.5 þ 4.2
deoxyribose phosphate backbone ‘tether’ are reflected in
Acyl phosphates
changes in the P–O ester (as well as other) torsional angles.
Acetyl phosphate 1.5 þ 4.8
These sequence-specific variations in the P–O (and C–O) tor- Carbamyl phosphate 1.1 þ 4.9
sional angles may explain the sequence-specific variations in
the 31P chemical shifts. a
All chemical shifts are reported with respect to an external 85% H3PO4 standard;
upfield shifts are given a negative sign.
b
Titrability: þ indicates that changes are observed in the chemical shift on changes in
31 pH: for phosphomonoesters this change is 4 ppm; for phosphoramidates 2.5 ppm; for
P NMR of Protein Complexes acyl phosphates 5.1 ppm; – indicates no change observed.

31
P NMR spectroscopy has proven to be very useful in the study
of various protein complexes. Table 4 provides an indication of Thus the pH dependence of the 31P signal of various phosphate
the range of 31P chemical shifts and the titration behaviour of monoesters bound to proteins can provide information on the
various phosphoprotein model compounds. Two examples of ionization state of the bound phosphate ester. For example,
such studies are described below. pyrimidine nucleotides, both free in solution and when bound
to bovine pancreatic ribonuclease A (RNase A), demonstrate
this point. The 31P chemical shift of free solution cytidine
Ribonuclease A
30 -monophosphate (30 -CMP) follows a simple titration curve,
Secondary ionization of a phosphate monoester produces and the ionization constant derived form the 31P shift variation
approximately a 4 ppm down field shift of the 31P signal. agrees with potentiometric titration values. The 31P chemical
NMR Spectroscopy, 31P 301

shift titration curve for the 30 -CMPRNase A complex, however, experiment (NOESY) described by Ernst and co-workers
cannot be analysed in terms of a single ionization process. Two involves three 90 pulses. Nuclei are frequency labelled by a
inflections observed in this titration indicated two ionizations variable delay time (t1) separating the first and second pulses.
with pK1¼4.7 and pK2¼6.7. The mixing time is between the second and third pulses, and
These results suggest that the nucleotide binds at around the detection of transverse magnetization as a function of time
neutral pH in the dianionic ionization state. Thus the (t2) follows the third pulse. During the mixing time, nuclei
30 -CMPRNase A complex 31P resonance is shifted upfield less labelled in t1 with a frequency corresponding to one site are
than 0.3 ppm from the free 30 -CMP between pH 6.5 and 7.5, converted by the exchange processes to a second site and evolve
whereas monoprotonation of the free dianion results in a in t2 with the frequency of the second site, giving rise to cross-
4 ppm upfield shift. Furthermore, the addition of the first peaks in the 2D spectrum. A 2D 31P exchange spectrum of PGM
proton to the nucleotide complex (pK2¼6.0–6.7) must occur shows cross-peaks indicating exchange between bound Glc6P
mainly on some site other than the dianionic phosphate and free Glc6P and between the two bound phosphorus sites,
because the 31P signal is shifted upfield by only 1–2 ppm. indicating transfer through free EP involving a full catalytic
The addition of a second proton (pK1¼4.0–5.7) to the com- cycle (see eqn [8]).
plex shifts the 31P signal further upfield so that at the lowest pH
values, the phosphate finally appears to be in the monoanionic
ionization state. Medical Applications of 31P NMR
On the basis of X-ray and 1H NMR studies, it is known that
the nucleotides are located in a highly basic active site with In vivo31P NMR spectroscopy and 31P magnetic resonance
protonated groups histidine-119, histidine-12 and, probably, imaging are also important applications of this nucleus. 31P
lysine-41, quite close to the phosphate. This suggests that pK1 signals from inorganic phosphate, adenosine triphosphate,
is associated with ionization of a protonated histidine residue adenosine diphosphate, creatine phosphate and sugar phos-
which hydrogen bonds to the phosphate. This highly positive phates can be observed in whole-cell preparations, intact tis-
active site, which is capable of perturbing the pK of the phos- sues, and whole bodies and can provide information on the
phate from 6 to 4.7, must have one or more hydrogen bonds to viability of the cells and tumour localization. Low sensitivity
the phosphate over the entire pH region. Yet at the pH extrema, continues to be a problem in widespread application of these
little if any perturbation of the 31P chemical shift is found. techniques. Additional details can be found in several of the
Apparently, the 31P chemical shift of the phosphate esters is entries in the Further Reading section.
largely affected by the protonation state and not by the highly
positive local environment of the enzyme.
Conclusions
31
P NMR has become an indispensable tool in studying the
31 chemistry and reactivity of phosphorus compounds, as well as
Two-Dimensional Exchange P NMR of
Phosphoglucomutase in studying numerous biochemical and biomedical problems.
Newer NMR instrumentation has enormously enhanced the
Phosphoglucomutase (PGM) catalyses the interconversion of sensitivity of the experiment and allowed 2D NMR studies to
glucose 1-phosphate and glucose 6-phosphate. The enzyme provide new means of signal assignment and analysis. Through
has 561 residues on a single polypeptide chain with molecular 2D and 3D heteronuclear NMR experiments it is now possible
weight 61 600 Da. Catalysis proceeds via a glucose 1,6- to unambiguously assign the 31P signals of duplex oligonucle-
bisphosphate intermediate where the formation and break- otides and other phosphate esters. Both empirical and theoret-
down of this intermediate results from two phosphate transfer ical correlations between measured coupling constants, 31P
steps involving a single enzymic phosphorylation site, Ser-116. chemical shifts, and structural parameters have provided an
A metal ion is required for activity and the most efficient metal important probe of the conformation and dynamics of nucleic
ion is the physiological activator, Mg2þ. The phosphate transfer acids, protein complexes, and small organophosphorus
steps are shown below. compounds.
k1 k6
EP þ GlclP ⇄ ED GlclP 6P ⇄ EP∗ þ Glc6P [8]
k1 k6
See also: In Vivo 1H MRS Methods; In Vivo NMR, Applications, Other
k1 k6
Ep þ GlclP ⇄ ED GlclP6P ⇄ EP þ Glc6P ∗
[9] Nuclei; Membranes Studied by NMR Spectroscopy.
k1 k6

EP and ED are the phospho and dephospho forms of the


enzyme, respectively, Glc1P is glucose 1-phosphate and
Glc6P is glucose 6-phosphate. Further Reading
Metal-free PGM and complexes with a variety of metal ions, Burt CT (1987) Phosphorus NMR in Biology, pp. 1–236. Boca Raton: FL CRC Press.
substrates and substrate analogues have been studied by 31P Crutchfield MM, Dungan CH, Letcher LH, Mark V, and Van Wazer JR (1967) Topics in
NMR. Under conditions where the enzyme is inactive, each of phosphorus chemistry. In: Grayson M and Griffin EF (eds.) Topics in Phosphorous
the three enzyme-bound intermediates in the above scheme Chemistry, pp. 1–487. New York: Wiley (Interscience).
Gorenstein DG (1984) Phosphorus-31 NMR: Principles and Applications, pp. 1–604.
can be studied. Orlando, FL: Academic Press.
Exchange processes can be detected by 2D 31P NMR in Gorenstein DG (1992) Advances in P-31 NMR. In: Engel R (ed.) Handbook of
addition to the conventional 1D methods. The 2D exchange Organophosphorus Chemistry, pp. 435–482. New York: Marcel Dekker.
302 NMR Spectroscopy, 31P

Gorenstein DG (1994) Conformation and dynamics of DNA and protein–DNA Quin LD and Verkade JG (1994) Phosphorus-31 NMR Spectral Properties in
complexes by 31P NMR. Chemical Reviews 94: 1315–1338. Compound Characterization and Structural Analysis, p. 1. New York: VCH.
Gorenstein DG (1996) Nucleic acids: Phosphorus-31 NMR. In: Grant DM and Harris RK Tebby JC (1991) Handbook of Phosphorus-31 Nuclear Magnetic Resonance Data, p. 1.
(eds.) Encyclopedia of Nuclear Magnetic Resonance, pp. 3340–3346. Chichester: Boca Raton, FL: CRC Press.
Wiley. Verkade JG and Quin LD (1987) Phosphorus-31 NMR Spectroscopy in Stereochemical
Karaghiosoff K (1996) Phosphorus-31 NMR. In: Grant DM and Harris RK (eds.) Analysis; Organic Compounds and Metal Complexes, pp. 1–455. Deerfield Beach,
Encyclopedia of Nuclear Magnetic Resonance, pp. 3612–3618. Chichester: Wiley. FL: VCH.
Mavel G (1973) Annual Reports on NMR Spectroscopy 5B: 1–350.

You might also like