Vous êtes sur la page 1sur 64

Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

www.elsevier.com/locate/pnmrs

Practical aspects of 51V and 93Nb solid-state NMR spectroscopy


and applications to oxide materials
O.B. Lapina
b

a,*

, D.F. Khabibulin a, A.A. Shubin a, V.V. Terskikh

a
Boreskov Institute of Catalysis, Prosp. Lavrentieva 5, Novosibirsk 630090, Russia
Steacie Institute for Molecular Sciences, National Research Council Canada, Ottawa, Ont., Canada K1A 0R6

Received 29 September 2007; accepted 6 December 2007


Available online 4 March 2008

Keywords: Solid-state NMR; Modern NMR techniques; NMR in catalysis; Vanadium oxide catalysts; Niobium oxide catalysts; Vanadia; Niobia; NMR
of Group VB elements; 51V NMR; 93Nb NMR; 181Ta NMR

Contents
1.
2.

3.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Group VB elements, NMR properties, and solid-state NMR concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1. Definition of NMR parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2. Vanadium-51 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3. Niobium-93 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4. Tantalum-181 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Modern NMR techniques most suitable for studying 51V and 93Nb in solids. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1. NMR spectra of stationary samples (quadrupolar echoes, QCPMG and nutations) . . . . . . . . . . . . . . . . . . . . . . . .
3.2. Magic-angle spinning (MAS) and high-speed magic-angle spinning (HS MAS) . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3. Satellite transition spectroscopy (SATRAS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4. Spinning sidebands analysis of selected transitions (SSTMAS). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5. MAS and static spectra analysis (MASSA) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.6. Multi-quantum MAS (MQMAS) and satellite transition MAS (STMAS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.7. Heteronuclear correlation spectroscopy (HETCOR) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.8. Double resonance experiments (SEDOR, REDOR, TRAPDOR, and REAPDOR) . . . . . . . . . . . . . . . . . . . . . . . .

129
131
131
131
133
134
134
134
135
136
139
139
141
144
144

Abbreviations: CP, cross polarization; CS, chemical shielding; CSA, chemical shift anisotropy; CT, central transition; CW, continuous wave; CW-NMR,
continuous wave nuclear magnetic resonance; DAS, dynamic-angle spinning; DeNOx, nitrogen oxide abatement; DFT, density functional theory; DOR,
double rotation; DQ, double-quantum; DQ STMAS, double-quantum satellite transition magic-angle spinning; EDAX, energy dispersive X-ray spectroscopy; EFG, electric eld gradient; ESR, electron spin resonance; FID, free induction decay; FT-NMR, Fourier transform nuclear magnetic resonance;
GIPAW, gauge-including projected augmented-wave; HETCOR, heteronuclear correlation spectroscopy; HFI, hyperne interaction tensor; HFMAS,
high-eld magic-angle spinning; HREM, high resolution electron microscopy; HSMAS, high-speed magic-angle spinning; KTN, KTa(1x)NbxO3; MAS,
magic-angle spinning; MASSA, magic-angle spinning and static spectra analysis; MQMAS, multiple quantum magic-angle spinning; MQ, multiple
quantum; NMR, nuclear magnetic resonance; NQR, nuclear quadrupolar resonance; QCPMG, quadrupolar Carr-Purcell Meiboom-Gill; PBN, Basubstituted Pb(Mg1/3Nb2/3)O3; PMN, Pb(Mg1/3Nb2/3)O3; PMN/PT, (1x)Pb(Mg1/3Nb2/3)O3/xPbTiO3; PSN, Sc-substituted Pb(Mg1/3Nb2/3)O3; PZN, Zrsubstituted Pb(Mg1/3Nb2/3)O3; REAPDOR, rotational echo adiabatic passage double resonance; REDOR, rotational echo double resonance; RF, radio
frequency; SATRAS, satellite transition spectroscopy; SBV, Strongly bound vanadium; SEDOR, spinecho double resonance; ST, satellite transition;
STMAS, satellite transition magic-angle spinning; SSTMAS, spinning sidebands analysis of selected transitions; TRAPDOR, transfer of population in
double resonance; VOCS, variable oset cumulative spectrum.
*
Corresponding author.
E-mail address: olga@catalysis.ru (O.B. Lapina).
0079-6565/$ - see front matter  2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.pnmrs.2007.12.001

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

3.9. Triple Resonance experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


3.10. Advantages of high magnetic field strengths. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. DFT and other quantum chemical computational approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5. 51V NMR data compilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1. Chemical shielding and quadrupolar tensor parameters in individual vanadium compounds. . . . . . . . . . . . . . . . . .
5.1.1. Tetrahedral Q0 sites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.2. Tetrahedral Q1 sites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.3. Tetrahedral Q2 sites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.4. Associated non-axial VO5 and VO6 sites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.5. Isolated and associated trigonal VO4 pyramids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.6. Isolated octahedral VO6 and tetragonal VO5 pyramids. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.7. Associated tetragonal pyramids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.8. Strongly associated octahedral sites in decavanadates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2. Correlating local environment of vanadium nuclei in VOx species with 51V NMR parameters . . . . . . . . . . . . . . . .
6. 93Nb NMR data compilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1. Chemical shielding and quadrupolar tensor parameters in individual niobium compounds . . . . . . . . . . . . . . . . . . .
6.1.1. Six-coordinated compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1.2. Four-coordinated compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1.3. Five-coordinated compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.1.4. Seven- and eight-coordinated compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2. 93Nb NMR chemical shift scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7. 181Ta NMR data compilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8. Paramagnetic effects in 51V and 93Nb solid-state NMR spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.1. Presence of paramagnetic cations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.2. Systems with vanadium in mixed oxidation states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9. Recent applications of solid-state 51V and 93Nb NMR in oxide materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.1. Applications of solid-state 51V NMR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.1.1. Bio-structural chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.1.2. Materials chemistry. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.1.3. Catalysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.2. Applications of solid-state 93Nb NMR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.2.1. Characterization of ferroelectrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.2.2. Silicates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.2.3. Miscellaneous applications. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.3. Multinuclear solid-state NMR in vanadia and niobia catalysts supported on Al2O3 . . . . . . . . . . . . . . . . . . . . . . .
9.3.1. Vanadia sites in VOx/Al2O3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.3.2. Niobia sites in NbOx/Al2O3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.3.3. Niobiavanadia species in (NbV)Ox/Al2O3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
10. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
There are many similarities between the rst three elements of the Group VB. Two of these three elements, vanadium and niobium, were discovered in 1801, and Tantalum
was discovered shortly after, in 1802. All three elements
bear mythological names: vanadium is named after the
Scandinavian goddess Vanadis because of its beautiful
multicolored compounds, and niobium is named after
Niobe, the daughter of Tantalus, the namesake of
Tantalum.
All three elements nd numerous applications in chemical industry, electronics, and metallurgy. Vanadium is a
very versatile metal: besides its main use in steel manufacturing, vanadia-based catalysts are frequently used for
large-scale sulfuric acid production [1], for cleaning ue

129

146
146
148
149
149
150
150
157
157
157
158
160
160
160
162
162
162
165
166
168
169
169
169
171
172
175
175
175
175
176
181
182
183
183
184
184
185
186
187
187
188

gases, for selective oxidation of hydrocarbons [24], for


reduction of nitrogen oxides with ammonia [5,6], and for
production of bulk chemicals [4,79]. Vanadium haloperoxidases [10] have potential as catalysts in industrial-scale
bio-catalytic conversions. Recent years have also brought
growing interest in niobium-based oxide systems often
showing improved catalytic properties [1113]. Even small
amounts of niobium oxide added to a catalytic mixture
may considerably enhance catalytic activity, selectivity,
and long-term stability [13]. Niobium oxide itself or mixed
with other oxides (Nb2O5SiO2, Nb2O5Al2O3, Nb2O5
TiO2, Nb2O5V2O5, etc.) is frequently used as a support
for catalytically active metals or other metal-oxide catalysts
[1113]. Hydrated niobium pentoxide (niobic acid,
Nb2O5nH2O) and niobium phosphate have been shown
to have unusually high surface acidity, signicant catalytic

130

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

activity, exceptional selectivity, and high stability in many


acid-catalyzed reactions [13].
From the NMR point of view, all three elements, vanadium, niobium, and tantalum, are quite similar as well.
Each element has only one dominant isotope suitable for
NMR spectroscopy, 51V, 93Nb, and 181Ta, each is quadrupolar having a half-integer nuclear spin.
Vanadium-51 (natural abundance 99.76%) has spin 7/2
and an electric quadrupolar moment of only 0.05 barn.
The relative receptivity of 51V NMR is 0.38 compared to
1
H NMR. The 93Nb nucleus (natural abundance 100%)
has spin 9/2. Niobium-93 is one of the most NMR-receptive nuclei with a receptivity of 0.482 relative to 1H and a
favorably low quadrupolar moment of 0.32 barn. On the
other hand, tantalum-181 (natural abundance 99%, spin
7/2) has one of the largest known quadrupolar moments,
3.44 barn, which hampers considerably 181Ta NMR experiments on solid samples (see below).
The rst solid-state 51V, 93Nb, and 181Ta NMR works
appeared in the late 1950s. The rst reported 51V NMR
spectra were recorded by Knight and Cohen in 1949 [14]
on polycrystalline Pb(VO3)3 and V2O5 samples to determine the magnetogyric ratio for the 51V nucleus. Similar
results for 93Nb were published in 1951 [15] and for 181Ta
in 1959 [16]. In 1961, Ragle and coworkers [17] described
the 51V magnetic anisotropy in a polycrystalline V2O5
sample.
In 19671969 several reports were published on singlecrystal 51V and 93Nb NMR. Gornostansky et al. determined parameters of the magnetic shielding anisotropy
and the quadrupolar coupling constants for 51V in V2O5
[18] and in KVO3 [19]. Single-crystal 93Nb NMR experiments had allowed one to measure the 93Nb quadrupolar
coupling constant in LiNbO3 [20,21].
Continuous wave (CW) NMR was successful in studying quadrupolar coupling parameters in the solid state,
since such spectra often showed well-dened discontinuities
corresponding to quadrupolar interactions (of the rstorder for a small value and of the second-order for a larger
value of quadrupolar constant). This approach was used to
determine quadrupolar coupling constants in several vanadates [18,2226]. Similar 93Nb experiments were performed
on LiNbO3 [2731].
With the development of pulsed FT-NMR instruments
it became more practical to determine magnetic shielding
parameters of 51V nucleus by NMR spectroscopy [32,33].
However, in these rst experiments the rst-order quadrupolar eects were often very dicult to observe.
Early 51V NMR results obtained with CW-NMR and
FT-NMR have been summarized by Pletnev et al. [22].
The introduction of sample spinning at the magic angle
to the external magnetic eld (MAS), with spinning speeds
up to 6 kHz, brought considerable improvements in spectral resolution allowing the identication of as many as
two or three non-equivalent vanadium sites in some compounds, and also allowed determination of the isotropic
chemical shift values for each site [34].

Unlike the early success of 51V NMR spectroscopy,


Nb NMR studies were somewhat slow to follow. In many
respects, this was because of the larger 93Nb quadrupolar
moment. In a solid the electric eld gradients arising from
the electronic cloud at the nucleus can interact with the
nuclear quadrupolar giving rise to considerable spectral
broadening. Nevertheless, several Nb-containing systems
were studied in detail using combined static and conventional MAS 93Nb NMR [3543].
Not surprisingly, 51V solid-state NMR has become an
important tool in characterizing the local structure of
vanadium sites in many vanadium-based systems [34,44].
Modern NMR techniques such as ultrahigh-speed MAS
with spinning speeds in excess of 35 kHz, MQMAS,
SATRAS, and others methods, have allowed accurate
information to be obtained on the local structure of vanadium sites, i.e. (i) the number of non-equivalent vanadium
sites, (ii) coordination numbers, (iii) the nature of atoms
in the rst coordination sphere, (iv) distortion of the rst
coordination sphere, (v) association of vanadiumoxygen
polyhedra. In addition, spinecho mapping spectra or
ultrahigh-speed MAS experiments have helped to identify
V5+ atoms bound via an oxygen atom to V4+ or another
paramagnetic species [45]. Defects and distortions in the
crystal structure can be revealed by analysis of distributions of the chemical shielding and quadrupolar tensor
parameters [46].
A growing interest in niobium-based catalytic systems,
as well as the practical importance of some Nb-containing
piezoelectric and optoelectronic materials, have stimulated
several recent 93Nb NMR studies employing modern solidstate NMR techniques [4752]. Using conventional and
ultrahigh magnetic eld facilities, ultrahigh-speed MAS,
DQ-STMAS, solid-echo and computer modeling, chemical
shifts and quadrupolar tensor parameters have been
reported for a considerable number of Nb compounds
[5355]. It has been shown, that the 93Nb isotropic shift
is sensitive to the coordination number of Nb sites. A
recently proposed 93Nb NMR chemical shift approach
allows determination of the coordination number in NbOx
polyhedra [55].
Because of the relatively large 181Ta quadrupolar
moment and the low resonance frequency, there is only a
very limited number of reports on 181Ta NMR in the solid
state [56]. Starting from the rst 181Ta NMR work on
KTaO3 [16], the total number of 181Ta NMR papers published so far is less than 10 [16,5763].
This review on the current state of the solid-state 51V
and 93Nb NMR spectroscopy includes both previously
published and original results. Advantages and restrictions
of various solid-state NMR techniques as applied to vanadium and niobium are discussed with illustrations from a
variety of vanadium and niobium containing oxide materials, including individual highly crystalline compounds,
solid solutions, glasses, and catalysts. The main purpose
of this work is to provide readers with the latest comprehensive compilation of 51V and 93Nb NMR data in oxide
93

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

materials, and to demonstrate the great potential of


and 93Nb NMR in solid-state chemistry.

51

the asymmetry parameter of the EFG (or nuclear quadrupolar) tensor.


For a chemical shift tensor we use the following denition for the isotropic chemical shift, diso, the chemical shift
anisotropy (CSA), dr, and the CSA asymmetry parameter,
gr :

2. Group VB elements, NMR properties, and solid-state


NMR concepts
The Group VB has only three nuclei most commonly
considered for NMR spectroscopy, 51V, 93Nb, and 181Ta.
The energy level diagrams for these three quadrupolar
nuclei are rather similar. An example for a spin-9/2 system,
i.e. for 93Nb, is shown in Fig. 1 together with the corresponding calculated powder patterns perturbed by the
rst-order and the second-order quadrupolar interactions.
NMR properties of 51V, 93Nb, and 181Ta are summarized
in Table 1.

gQ

dr dZZ  diso
dYY  dXX
gr
dZZ  diso

3
4

^d ^Iriso  r
^

V YY  V XX
V ZZ

Here, ^I is the unit matrix and riso is the isotropic value of


the chemical shielding tensor for a selected reference compound. Absolute chemical shieldings for some commonly
used reference compounds can be found in [64].

In this review, we will use the following convention for


the quadrupolar coupling constant, CQ (MHz), and the
asymmetry parameter, gQ, as:
eQV ZZ
;
h

diso 13dXX dYY dZZ

Here, dXX, dYY, and dZZ are the principal components of


the CSA tensor.
Note that the CSA tensor ^d is related to the chemical
^ as
shielding anisotropy tensor r

2.1. Denition of NMR parameters

CQ

131

2.2. Vanadium-51
The vanadium-51 nucleus has high NMR receptivity
and a convenient resonance frequency, which is very close
to that of the frequency of 13C. The 51V NMR spectra of
solid samples show not only quadrupolar interactions,
but also sizable magnetic shielding eects. When present

Here, Q is the quadrupolar moment of the nucleus, and


VXX, VYY, VZZ are the principal values of the traceless electric eld gradient (EFG) tensor Vij ordered in a sequence
|VZZ| P |VXX| P jVYY| with the common designation
VZZ = eq for its largest principal component, while gQ is

I-order quadrupolar
B0 only
hCQ
Zeeman Interaction
(3 cos 2 1)
288
-h0m

96 0
9(16sin + 2sin 2 )
4

II-order quadrupolar
hCQ

-9/2
2

7(0sin + 6sin 2 )

-1

5(9sin 2 - 12sin )

-3

3(11sin 2 - 20sin )

-15/4

(12sin 2 - 24sin )

15/4

(24sin - 12sin 2 )

-7/2
-5/2
-3/2
-1/2
1/2

3(20sin - 11sin 2 )
4

3/2

-3

5(12sin - 9sin 2 )
4

-1
2
6

5/2

400

200

-200

-400

7/2

-7(0sin + 6sin 2 )
4

-9(16sin + 2sin 2 )
4

9/2

20000

10000

-10000

-20000

(ppm)
Fig. 1. Eects of the quadrupolar interactions on a spin-9/2 system in a solid sample. (A) Energy level diagram. (B) Calculated rst-order powder pattern
at 9.4 T including all transitions. (C) Second-order powder pattern of the central transition for gQ = 0. Simulation parameters CQ = 20 MHz, gQ = 0,
m0 = 97.9 MHz.

132

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

Table 1
NMR properties of the Group VB elements
Isotope

51

V
Nb
181
Ta
93

Spin

Natural
abundance (%)

7/2
9/2
7/2

Quadrupolar
moment (barn)

0.05
0.32
3.44

99.75
100.0
99.99

0.38
0.48
0.04

together, these two contributions can considerably complicate NMR spectra and their analysis. When studying
multiphase amorphous systems, such as most of the vanadia-based heterogeneous catalysts, the 51V NMR spectra
become extremely complicated, and advanced NMR techniques together with spectral simulations are often required
for their successful interpretation.
It is common for 51V to exhibit quadrupolar coupling
constants ranging from 2 to 6 MHz, but rarely exceeding
10 MHz [34]. At the same time the magnetic shielding
anisotropy is normally below 1000 ppm, and is often found
within the 100500 ppm range, depending on the coordination environment.
Simulated static spectra of all NMR transitions for a
spin-7/2 nucleus, i.e. 51V, with a quadrupolar coupling constant CQ = 4.5 MHz are shown in Fig. 2A (CQ = 4.5 MHz,
gQ = 01). The whole spectrum, including the outer
5/2 M 7/2 transitions, extends over 1.5 MHz, which
requires a radio-frequency pulse for its homogeneous excitation shorter than a few tenths of microsecond.
Modern FT-NMR solid-state spectrometers are wellequipped for obtaining 51V NMR spectra in solid materials

NMR frequency
(MHz) at

Sensitivity
relative to 1H

9.4 T

21.1 T

105.2
97.8
47.9

236.6
220.0
107.7

Chemical shift
range (ppm)

Reference
sample

2000
4000
3450

VOCl3
NbCl5/CH3CN
K[TaCl6]

with all the quadrupolar satellite transitions present. Minimal requirements are reasonably high magnetic elds
(P9 T), MAS spinning speeds exceeding 15 kHz, extremely
short radio frequency pulses, normally shorter than 0.5 ls,
and very fast digitizing rates. For example, a 20 MHz
(50 ns) digitizing rate is necessary to provide a 10 MHz full
spectral width.
Along with the quadrupolar interactions, as already
mentioned above, it is also important to consider the magnetic shielding while analyzing 51V NMR spectra. The
magnetic shielding is commonly described as a chemical
shift anisotropy tensor with three principal tensor components. The chemical shift anisotropy is most pronounced
in the central transition. This is illustrated in Fig. 2B for
a series of spectra calculated with dierent CSA asymmetry
parameters, gr.
For 51V NMR the central transition is very often
aected by the rst-order quadrupolar interactions
(Fig. 3). Both the quadrupolar and the magnetic shielding
interactions are described by second-rank tensors, which
make it necessary to dene their relative orientation. This
relative orientation is given by three Euler angles, a, b,

1.0

1.0

0.8
0.75

0.6
0.5
0.4

0.25
0.2

0.0

0.0

600

400

200

-200

, ppm

-400

-600

1000

500

-500

-1000

, ppm

Fig. 2. NMR powder patterns calculated for a spin-7/2 system of a solid polycrystalline sample at 105.2 MHz. (A) The rst-order quadrupolar eects at
dierent values of gQ (CQ = 4.5 MHz). (B) Chemical shielding anisotropy eects on the central transition at dierent values of gr (dr = 500 ppm, CQ = 0).

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

133

1
1
2
2

3
4

4
5
6
7
8
500

-500 -1000 -1500 -2000

5
500

-500 -1000 -1500 -2000

, ppm

Fig. 3. Combined eects of the quadrupolar and magnetic shielding interactions on the NMR spectra for a spin-7/2 system of a solid, polycrystalline
sample. (A) Simulated with NMR parameters typical for V2O5, m0 = 105.2 MHz CQ = 0.797 MHz, gQ = 0.08, diso = 609 ppm, dr = 645 ppm, gr = 0.11,
a = 42, b = 126, c = 5; (1) full spectrum and transitions (2) +7/2 M +5/2, (3) +5/2 M +3/2, (4) +3/2 M +1/2, (5)+1/2 M 1/2, (6) 1/2 M 3/2, (7)
3/2 M 5/2, (8) 5/2 M 7/2. (B) Eects of the relative orientation of the CS and quadrupolar tensors. Only the Euler angle b is varied as following: (1)
b = 160, (2) b = 140, (3) b = 126, (4) b = 110, (5) b = 90.

and c. Therefore, in general case, there are eight independent parameters in the Spin-Hamiltonian describing a
solid-state 51V NMR spectrum. These parameters are three
principal components of the CSA tensor, any two principal
components of traceless quadrupolar tensor, and three
Euler angles describing relative orientation of the CSA
and quadrupolar tensors (see also Eqs. (1)(4) above for
the denition of commonly used corresponding NMR +
parameters).
In the rst-order of perturbation theory, each NMR
transition has only three singularities (Fig. 3). It is therefore necessary for spectra of static samples to analyze at
least three separate transitions in order to obtain the full
set of eight independent parameters. This requires observation of not only the central +1/2 M 1/2 transition, but
also two satellite transitions. The central transition lineshape can be used to determine the CSA parameters, while
the quadrupolar interaction parameters can then be determined from the satellite transitions. We note, however, that
the magnetic shielding aects not only the central transition but the satellite transitions as well, i.e. the singularities
are now aected by both interactions (Fig. 3B).
Let us consider the most narrow 3/2 M 1/2 satellite
transitions in the magnetic eld of 9.4 T and the quadrupolar coupling constant of CQ = 4.5 MHz (gQ = 0). The total
width of these transitions will be about 4500 ppm due to
the quadrupolar interactions only. These transitions are
wider by some 15% if the magnetic shielding, dr = 500 ppm
and gr = 0, is also present. In this case if the magnetic
shielding eects are not taken into consideration, the calcu-

lated quadrupolar coupling constant could be easily overestimated by as much as 15%.


2.3. Niobium-93
93

Nb NMR spectra, even at very high magnetic elds,


are often dominated by the quadrupolar interactions. For
the central transition +1/2 M 1/2, the second-order
quadrupolar perturbation results in a characteristic powder
pattern. Non-central, or satellite, transitions are spread far
from the central Larmor frequency. At the same time, for
I = 9/2 nuclei, the satellite transitions are closer to the central transition compared with those from nuclei with lower
spin quantum numbers (for a given CQ). Thus, for I = 9/2
not only a central transition, but also several satellite transitions can routinely be observed using conventional solidstate NMR spectrometers.
At present, there is no universal method for quadrupolar nuclei with a half-integer spin that allows one to obtain
a complete set of the quadrupolar and chemical shielding
(CS) tensor parameters for the wide range of values of
the quadrupolar constant and the chemical shift anisotropy
often found for 93Nb. Depending on the absolute values of
these parameters and on their relative magnitude only certain NMR techniques or a combination of several NMR
techniques can be applied successfully. As it has been suggested in [4754,6567], the most suitable techniques for
93
Nb are experiments at high magnetic elds applying
high-speed MAS, solid-echo, MQMAS, pure-phase nutation, and STMAS techniques [55].

134

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

2.4. Tantalum-181
Due to the relatively large 181Ta quadrupolar moment
(3.44 barn), 181Ta NMR studies on solid samples are rare
and seriously hampered by strong line broadening caused
by quadrupolar relaxation. Another complication is the
low resonance frequency and the consequent low receptivity. 181Ta may be considered as a low-gamma nucleus and
as such is very dicult to work with, particularly for solid
samples. However, the availability of ultrahigh magnetic
elds for solid-state NMR research makes 181Ta NMR feasible, yet challenging. The NMR spectra of 181Ta for solid
samples are similar to those of 51V and 93Nb and will be
represented by a set of central and satellite transitions,
unless the nucleus is in an environment with high symmetry. The large 181Ta quadrupolar moment ensures that
the NMR spectra will most likely be dominated by quadrupolar interactions, but magnetic shielding eects should not
be ignored.
3. Modern NMR techniques most suitable for studying
and 93Nb in solids

51

3.1. NMR spectra of stationary samples (quadrupolar


echoes, QCPMG and nutations)
The rst solid-state 51V and 93Nb NMR spectra were
recorded on stationary samples [22,24,2732,6870]. Even
though numerous line-narrowing techniques like MAS
and MQMAS have since been introduced, recording
NMR spectra on static samples remains important and
often indispensable for proper interpretation of the NMR
data either in individual crystalline compounds or in such
complex multiphase systems as heterogeneous catalysts.
In most cases, static 51V and 93Nb NMR spectra represent a superposition of the central and satellite transitions.
By manipulating with the excitation bandwidth of rf pulses,
it is possible to obtain separately the spectrum of the central transition only, and the spectrum including all or at
least several satellite transitions. However, increasing the
sweep width in order to record broader spectra often leads
to amplied acoustic ringing of the probehead. When
acoustic ringing is present, it causes the loss of the rst
few points in the FID, which results in signicantly distorted spectra. In the worst cases, the very broad spectra
can be lost altogether solely due to the dead time of the
probe/receiver. This problem can be circumvented by
applying techniques involving spinecho pulse sequences.
There are two main types of echo pulse sequences used
in solid-state NMR of quadrupolar nuclei with a half-integer spin: the Hahn-echo [71] and the solid-echo [65,72]. In
the solid-echo (or quadrupolar echo) pulse sequence,
p1  s1  p2  s2  AQ, rf pulses p1 and p2 are not necessarily multiples of a p/2 pulse, which makes this sequence
useful for recording very broad spectra. In the Hahn-echo
pulse sequence the pulse durations are dened as
p2 = 2 p1 and are normally multiples of the p/2 pulse.

For NMR of qudrupolar nuclei, such as 51V and 93Nb,


with many satellite transitions present, it is often dicult
to dene a single p/2 pulse. This can be done, however,
for each separate transition, for example, for the central
transition only. In practice, the Hahn-echo pulse sequence
is often used not only to obtain the spectra of central transitions, but also to carefully calibrate the p/2 pulse, which
then can be applied in more advanced experiments employing various n-quantum coherence lters.
The line shape in static spectra, either of the central
transition, or when superimposed with several satellite
transitions, is very sensitive to durations of pulses in the
echo sequences and to delays between pulses. Inaccurate
choice of pulses or delays may lead to distorted line shapes,
and therefore to an incorrect interpretation of the spectra.
The situation becomes even more complicated, when there
is more than one site present in the system. The eects of
quadrupolar echoes for half-integer spins in static solidstate NMR spectra have recently been discussed in
[65,73,74].
Improved signal-to-noise ratio in static spinecho spectra can be achieved by recording the whole echo spectra,
via a modied solid-echo pulse sequence, p1  s1 
p2  AQ, as demonstrated by Massiot et al. [75] and Wu
and Dong [76]. The whole-echo technique is benecial
when the intrinsic spinspin relaxation time of the sample,
T2, is long suciently and the lines are suciently broad to
avoid truncation. In such cases the
p sensitivity can be
improved quite easily by a factor of 2, while still preserving the correct lineshape. The whole-echo acquisition is
often used to improve signal-to-noise in MQMAS and similar experiments (see below).
Recently re-introduced to NMR of half-integer quadrupolar nuclei, a Carr-Purcell Meiboom-Gill technique,
QCPMG, provides even greater sensitivity enhancement
over the conventional spinecho [7781]. In QCPMG
experiments, a standard solid-echo pulse sequence as
shown above is followed by a series of p pulses with a
whole-echo acquisition after each p pulse. The resulting
train of whole echoes can than be Fourier transformed to
give a series of equally spaced sharp spikelets outlining
the static powder pattern, somewhat resembling an MAS
spectrum. Another approach in processing QCPMG spectra involves adding together all echoes in the train and then
treating the resulting sum as a regular whole echo
spectrum.
An example of a 93Nb QCPMG NMR spectrum
recorded for La3NbO7 is shown in Fig. 4. In this case use
of the QCPMG spectrum reproduces very well the static
powder pattern obtained for the same sample via use of a
traditional spinecho approach. Either of the two spectra
can be used to determine the quadrupolar and CSA parameters for Nb sites in this compound. While the two spectra
are of about the same quality, equally suitable for further
analysis, the QCPMG spectrum required only 64 scans to
acquire, while the solid-echo spectrum was obtained in
4096 scans, i.e. took 64 times longer to acquire. Generally

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

-800

-1000

, ppm

-1200

-1400

Fig. 4. 93Nb NMR spectra of a stationary La3NbO7 sample of a


polycrystalline solid recorded at 21.1 T. (1) Experimental spinecho
spectrum. (2) Spectrum (1) simulated with the following parameters:
CQ = 49 MHz, gQ = 0.275, diso = 968 ppm, dr = 113 ppm, gr = 0.69,
a = 50, b = 27, c = 72. (3) Experimental QCPMG spectrum. The
number of accumulated scans was 4096 for (1) and 64 for (3).

speaking, a QCPMG pulse sequence containing n full echoes in its acquisition p


cycle delivers sensitivity enhancement
by a factor of n 2 over a traditional single-echo
experiment.
Because of this considerably improved sensitivity, today
the QCPMG technique nds many applications in solidstate NMR of low-gamma and low natural abundance
quadrupolar nuclei. Most recently QCPMG, often combined with a variable oset cumulative spectrum approach,
VOCS, is being used for systems with extremely large quadrupolar coupling constants [82]. Particularly large quadrupolar coupling constants are also expected for most
tantalum compounds. Naturally, in such systems 181Ta
QCPMG NMR should become the exploratory technique
of choice. 51V and 93Nb NMR benets from QCPMG in
cases where the vanadium or niobium content is small,
for example, in many supported catalysts, or in vanadiumor niobium-doped electronic materials.
While valuable, QCPMG does not always provide very
accurate line shapes as shown by Ooms et al. [83], and this
may complicate analysis of the spectra, or even lead to a
wrong interpretation of the experimental data. QCPMG
also requires suciently long T2 relaxation times to acquire
a train of echoes. Unfortunately in many important systems such as supported heterogeneous catalysts T2 relaxation times are often not long enough.
For stationary samples, Samoson and Lippmaa have
developed a nutation technique, which is based on the
acute sensitivity of the static line shape to the duration of
an excitation rf pulse [84]. By comparing the nutation frequency with the frequency of the excitation rf pulse it is
possible to determine the quadrupolar coupling parameters. The nutation technique is not very useful for 51V, since
for this nucleus it is typical to observe only the rst-order

135

quadrupolar interactions. In contrast, for the 93Nb NMR


with strong second-order quadrupolar eects even in the
highest magnetic elds, the nutation technique has been
proven quite informative, particularly when performed in
a two-dimensional fashion [48,50].
One of the advantages of recording static NMR spectra
of quadrupolar nuclei, including 51V and 93Nb, is simplicity
of execution, and that it does not involve purchasing or
building very expensive and maintenance-demanding
MAS probes. Also, the static line-shape does not require
signicant computing power to analyze. On the other hand,
static NMR spectra have limited use mostly due to their
broadness and the resulting insucient spectral resolution.
It is also a very challenging task to obtain correct line
shapes. When there are more than one or two individual
sites present, it often becomes impossible to interpret a
spectrum. Even the slightest distortion in the local nucleus
environment may render useless any attempts to obtain
meaningful spectral information.
3.2. Magic-angle spinning (MAS) and high-speed magicangle spinning (HS MAS)
The rst 51V MAS NMR spectra of vanadium compounds [85], were recorded by using MAS to successfully
minimize eects of dipolar interactions, magnetic shielding,
and the rst-order quadrupolar interactions. At the same
time, the low spinning speeds used, frequently below
5 kHz, often resulted in multiple spinning sidebands overlapping with isotropic 51V lines thus complicating the interpretation of the spectra. In these earlier 51V NMR works,
MAS sidebands were considered a nuisance, and every
attempt was made to minimize their number in the spectra.
Low MAS spinning speeds have little eect in resolving
93
Nb NMR spectra broadened by strong quadrupolar
eects.
Recent advances in the MAS probe technology have
resulted in considerably increased spinning speed rates
and improved spinning stability. All major manufacturers
of solid-state NMR equipment are now oering MAS
probes capable of spinning speeds ranging from 35 to
70 kHz [8690]. However, the nal choice of the MAS spinning speed still greatly depends on the system under investigation and on the applied NMR technique. For example,
the spinning speed should be as low as possible in order to
analyze the spinning sidebands of the satellite transitions
(SATRAS, see below). At the same time, high spinning
speeds or several dierent spinning speeds are often
required to resolve non-equivalent sites.
The 51V NMR chemical shift range in solid vanadiabased systems has been reported to exceed 1200 ppm, with
most of the shifts falling within ca. 500 ppm. To completely
free this spectral window from MAS spinning sidebands
while performing experiments at 9.4 T (51V resonance
frequency of 105.2 MHz), it would be necessary to spin
the sample at speeds exceeding 40 kHz. Lower spinning
speeds may render impossible a correct interpretation of

136

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

the spectra in a multi-component system. At higher magnetic elds, the MAS spinning speed should be even faster,
i.e. exceeding 60 kHz at 14 T, and 90 kHz at 21 T.
Spectral resolution in a MAS spectrum is limited by the
line width of the isotropic lines. In 51V NMR spectra the
rst-order quadrupolar eects and the magnetic shielding
eects are eectively averaged by fast MAS. At the same
time, the second-order quadrupolar eects are averaged
by MAS only partially, thus leading to a residual broadening of the lines even under innitely fast MAS. Additional
broadening in the spectra is introduced by a distribution of
the NMR parameters due to the nature of the sample, i.e.
due to various defects, surface eects, a lack of, or low level
of crystallinity, etc. In such cases a numerical simulation of
the spectrum is the only suitable approach to achieve a correct interpretation. Only a few of the currently available
computational tools for spectra simulations, oer the
option of including the distribution of NMR parameters
in their tting procedures [91].
The distribution of the magnetic shielding parameters
and the quadrupolar coupling parameters aect the
NMR spectra in a somewhat similar fashion. In addition
to overall line broadening, the singularities in the line shape
are less pronounced. It is important to note that MAS spectra are more sensitive to the distribution of the magnetic
shielding parameters (Fig. 5B, spectrum 2), while the distribution of the quadrupolar parameters is particularly obvious in the static spectra (Fig. 5A, spectrum 3). The
distribution in both, the magnetic shielding and the rstorder quadrupolar interactions, results in a homogeneous
line broadening, while the distribution in the second-order
quadrupolar interactions results in a quasi-homogeneous
line broadening.
In the case of 93Nb NMR, when the line width is mostly
governed by the quadrupolar interaction, only the highest
available MAS spinning speeds will be eective at low
magnetic elds. Considerably improved resolution in the
93
Nb NMR spectra can be achieved at the highest available

magnetic elds by applying the highest available MAS


spinning speeds.
Today, the high-speed MAS technique is frequently
applied in a combination with advanced line-narrowing
pulse sequences, including MQMAS, STMAS, and CP/
MAS.
3.3. Satellite transition spectroscopy (SATRAS)
The SATRAS technique is based on a numerical analysis of the integral intensities of the MAS spinning sidebands [9296]. Therefore, both the spectral resolution
and the spinning stability are important factors in the
SATRAS spectra analysis. As an example, some simulated
SATRAS spectra are shown in Fig. 6A. It is common to
observe a spectrum being a superposition of the several
transitions. The resulting spinning sideband patterns can
be quite complicated, which sometime makes it very dicult to interpret them (Fig. 6). For example, in the case
of 51V, every spinning sideband is often a superposition
of up to four individual lines from dierent satellite transitions. Because the spinning sidebands from dierent satellite transitions have dierent line width and slightly
shifted relative to each other, it is important to know what
contribution each makes into the integral intensity being
measured.
Eects of the line broadening on the MAS spinning sidebands of the transitions 7/2 M 5/2, 5/2 M 3/2, and
3/2 M 1/2 are shown in Fig. 6B. Even the slightest additional broadening, here by only 100 Hz, may lead to considerable broadening of the spinning sidebands from
some transitions, i.e. 7/2 M 5/2, making them impossible to detect by NMR. As a result, in some cases the integral intensities of the spinning sidebands cannot be
measured with sucient precision.
It is known [97], that the integral intensity of a transition
m M (m  1) is proportional to I(I + 1)  m(m  1), i.e.
for a spin-7/2 nucleus this means that the integral intensity

B
1

3
2000

1000

, ppm

-1000

-2000

2000

1000

-1000

-2000

, ppm

Fig. 5. Eects of distributions of NMR parameters on (A) powder patterns in stationary samples and (B) on 5 kHz MAS spectra. (1) Calculated for a spin7/2 system at 105.2 MHz without distributions with the following parameters, dr = 200 ppm, gr = 0, CQ = 2 MHz, gQ = 0.2. (2) As in (1) but with diso
distribution of 20 ppm. (3) As in (1) but with CQ distribution of 400 kHz.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

137

B
1
1
2

2
8000

4000

0
-4000
, ppm

3
-8000

1800

1500
, ppm

1200

Fig. 6. Simulated MAS NMR spectra for a spin-7/2 system. (A1) As-recorded SATRAS spectrum. (A2) Integral intensity spectrum of the spinning
sidebands. (B1 and B2) Eects of the broadening on the MAS spinning sidebands of the transitions 7/2 M 5/2, 5/2 M 3/2, and 3/2 M 1/2. (B1)
10 Hz broadening. (B2) 100 Hz broadening. (B3) Intensities of the satellites for all transitions (black stack) and without a 7/2 M 5/2 transition (gray
stack). Spectra were calculated at 105.2 MHz with the following parameters, CQ = 5 MHz, gQ = 0.4, dr = 200 ppm, gr = 0.7, a = 20, b = 20, c = 0,
mr = 10 kHz.

of the spectrum is distributed as 16:15:12:7 from the central


transition to the satellite transitions. If the 7/2 M 5/2
transition becomes invisible, this would result in the loss
of up to 14% of the total spectral intensity. However, the
contribution of each satellite transition into the spinning
sidebands depends upon the quadrupolar and magnetic
shielding parameters and the MAS spinning speed, i.e.
the resulting distribution of intensities across the spectrum
may not even follow the pattern mentioned above. All this
may lead to considerable errors in determining the spectral
parameters from the analysis of the SATRAS spectra.
To somewhat minimize possible errors, it is critical to
carefully choose the proper experimental conditions and
the pulse sequence. Selection of adequate rf pulse durations, rf power, pulse phases, relaxation delays, all help
to dene which part of the spectrum is being excited, i.e.
which satellite transitions are being observed.
According to Samoson [98], the spectral width, d(2)(m),
and broadening, D(m), caused by the second-order quadrupolar interactions for a particular satellite transition are
given by:
!
2
g2Q
3 C Q II 1  9mm  1  3
2
d m
1
6
40 m20
3
I 2 2I  12
!
2
g2Q
3 C Q 6II 1  34mm  1  13
Dm
1
7
128 m20
3
I 2 2I  12
where I is the nuclear spin, quantum number m = 1/2 represents the central transition, while m = 3/2 represents the
satellite transitions between m = 1/2 and m = 3/2, etc.
Relative spectral widths of the spinning sidebands from different transitions for a spin-7/2 nucleus are summarized in
Table 2. The broadest are the 7/2 M 5/2 transitions,
while all other satellite transitions are actually narrower
than the central transition. It is clear, that the transitions
+1/2 M 1/2, 1/2 M 3/2, and 3/2 M 5/2 are the easiest to observe experimentally. At the same time, recording

Table 2
Relative quadrupolar shift d(2)(m) and line broadening Dm of the spinning
sidebands of the satellite transitions caused by the second-order quadrupolar interactions for spin-7/2 nuclei
m

1/2

3/2

5/2

7/2

Dm/D(m = 1/2)
d(2)(m)/d(2)(m = 1/2)

1
1

0.622
0.4

0.511
1.4

2.4
4.4

the 5/2 M 7/2 transitions with sucient signal-to-noise


ratio would require much longer accumulation times.
Because the SATRAS technique employs analysis of the
integral intensities of the satellite transitions, it is more convenient to perform calculations taking into account only
+1/2 M 1/2, 1/2 M 3/2, and 3/2 M 5/2 transitions, and to exclude the 5/2 M 7/2 transitions. This
considerably simplies the analysis and normally produces
much more accurate results. The full MAS spectrum of all
four quadrupolar transitions is shown in Fig. 7. Unaided
analysis of such spectra is not straightforward, since even
visible singularities cannot unambiguously be assigned to
any particular satellite transition. The following algorithm
has been developed in our group to simplify analysis of the
SATRAS spectra of quadrupolar nuclei with half-integer
spin greater than 3/2, including 51V and 93Nb.
First, the individual satellite transitions should be identied in the spectrum (Fig. 7). Since the outermost satellite
transitions are the most sensitive to the quadrupolar interactions, it is convenient to use these transitions to estimate
the upper limit of the quadrupolar coupling constant.
Excluding the 5/2 M 7/2 transitions (Fig. 7, spectra 3
and 8) as mentioned above, the next outermost satellite
transitions, 3/2 M 5/2, should be analyzed (Fig. 7, spectra 4 and 7). These transitions have the narrowest spinning
sidebands, which makes it easier to identify these particular
transitions. In general, any satellite transition (m  1) M
m is asymmetric with the line shape depending on the

138

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

1000

0
, ppm

-1000

x10
x2

5
6

x2

7
8

10000

x10
5000

-5000

-10000

, ppm
Fig. 7. Simulated MAS NMR spectra for a spin-7/2 system. (1) Full spectrum representing a superposition of all transitions. (2) Central +1/2 M 1/2
transition. (38) Satellite transitions +7/2 M +5/2, +5/2 M +3/2, +3/2 M +1/2, 1/2 M 3/2, 3/2 M 5/2, 5/2 M 7/2, respectively. The vertical
intensities of some satellite transitions are scaled as shown. Spectra were calculated at 105.2 MHz with the following parameters, CQ = 5 MHz, gQ = 0.4,
dr = 200 ppm, gr = 0.7, a = 20, b = 20, c = 0, mr = 10 kHz.

Euler angles dening the relative orientations of the chemical shift and quadrupolar tensors. As a result the following
simplied analysis may lead to somewhat overestimated
values of the quadrupolar coupling constants.
Each satellite transition line shape has three singularities, similar to the three singularities produced by values
of the chemical shift tensor for the central transition. For
the transitions 3/2 M 5/2, these singularities are the
most pronounced and easiest to use in estimates. We note
that in the static spectra with a non-zero asymmetry
parameter of the quadrupolar interaction, the singularities
are practically unobservable. At the same time in the MAS
spectra these singularities can be seen as a characteristic
shape of the spinning sideband envelope. This becomes
particularly important for the 3/2 M 5/2 satellite transitions, when the intensity of the spinning sidebands
decreases sharply at the positions of singularities, and is
additionally highlighted by very broad spinning sidebands
from the 5/2 M 7/2 transitions (Fig. 8).

7500

7000

6500
, ppm

6000

5500

Fig. 8. Simulated MAS NMR spectrum for a spin-7/2 system showing


details of the 3/2 M 5/2 transition (narrow lines) and the 5/2 M 7/2
transition (broad lines). Spectra were calculated at 105.2 MHz with the
following parameters, CQ = 5 MHz, gQ = 0.4, dr = 200 ppm, gr = 0.7,
a = 20, b = 20, c = 0, mr = 10 kHz.

When the 3/2 M 5/2 transitions have been identied,


it is then possible to estimate the value of the quadrupolar
coupling constant. All other NMR parameters responsible
for the line shape of the satellite transitions, including the
asymmetry parameter and the Euler angles, cannot be estimated in a similar simplied fashion. To obtain these
parameters a complex numerical analysis of the integral

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

intensities of the spinning sidebands is required, often


involving an extremely time consuming minimization procedure and requires considerable computing resources [99].
To summarize, the SATRAS technique has a certain
advantage of being able to precisely determine all NMR
parameters at once from a single one-dimensional NMR
spectrum. At the same time this technique is applicable
and limited only to highly crystalline samples with small
to moderate quadrupolar coupling constants. Another
obvious limitation is the large volume of calculations
required.
3.4. Spinning sidebands analysis of selected transitions
(SSTMAS)
It is not uncommon while performing SATRAS analysis
to obtain not one but several dierent sets of NMR parameters tting the experimental spectrum equally well. In such
cases another analytical approach can be employed to identify the correct set. This approach uses lineshape analysis of
spinning sidebands from selected satellite transitions, thus
called spinning sidebands analysis of selected transitions
or SSTMAS. This approach has been found particularly
useful for spin-9/2 systems including 93Nb [54,55].
As a rst iteration step, an approximate value of the
quadrupolar coupling constant can be estimated from the
low-eld static spectra. This approximate CQ value is further rened via lineshape analysis of spinning sidebands
from selected satellite transitions in MAS spectra. In the
case of 93Nb, the analysis is usually performed for the satellite 3/2 M 5/2 transitions and the central +1/2 M 1/2
transition. Relative shifts and broadening of the satellite
transitions with respect to the central transition for I =
9/2 were calculated earlier by Du et al. [53,54] using Eqs.
(6) and (7) [98] (Table 3).
As shown in Table 3, for I = 9/2 the m = 5/2 satellite
transition has the smallest broadening of 0.055 in respect
to the central transition. As for the second-order quadrupolar shift, only for the m = 3/2 satellite transition the
quadrupolar shift has the same sign as the central transition. These two eects are important in calculating the
quadrupolar coupling constants from experimental 93Nb
MAS spectra.
The range of quadrupolar coupling constants that can
be estimated like this is 1550 MHz at 9.4 T and 20
80 MHz at 21.1 T [55]. In experimental spectra the lineshape of spinning sidebands will also depend on variations

Table 3
Relative quadrupolar shift d(2)(m) and line broadening Dm of the spinning
sidebands of the satellite transitions caused by the second-order quadrupolar interactions for spin-9/2 nuclei [54]
m

1/2

3/2

5/2

7/2

9/2

Dm/D(m = 1/2)
d(2)(m)/d(2)(m = 1/2)

1
1

0.764
0.625

0.055
0.5

1.125
2.375

2.778
5

139

in gQ [53,54]. A numerical simulation of the full spectrum


and the central transition of MAS spectra allows for accurate determination of CQ, gQ, and diso values.
Using the values of CQ, gQ, and diso obtained from the
analysis of satellites, it is then possible to calculate the
chemical shift anisotropy parameter dr from analysis of
the static spectra recorded at dierent magnetic elds.
The static spectra recorded at higher elds are more informative in determining the chemical shift tensor parameters.
For example at 21.1 T even the small values of dr
(100 ppm) can be calculated accurately for compounds
with CQ  20 MHz, as discussed below.
The SSTMAS technique can be illustrated with an
example of 51V NMR in LaVO4, where the second-order
quadrupolar interaction has to be taken into account.
Experimental 51V MAS spectra recorded for LaVO4 are
shown in Fig. 9. To perform a SATRAS analysis of this
spectrum it is necessary rst, to measure and plot integrated intensities from a hundred or so MAS spinning sidebands. Next, computer modeling needs to be performed to
t the experimentally measured intensities. In the SSTMAS
technique, just a single line with a pronounced quadrupolar
shape is often sucient for a complete analysis. As shown
in Fig. 9 by a dotted line, the t of the central line obtained
with the SSTMAS technique has been obtained by taking
into account not only the quadrupolar coupling and
the magnetic shielding tensors, but also their relative
orientation.
Also shown in Fig. 9 are the calculated spectra of the
central line with xed parameters of the quadrupolar and
chemical shielding tensors including their relative orientation. Acute sensitivity of the line shape of selected transitions in MAS spectra to variations in NMR parameters
makes it possible to apply this technique in cases when full
MAS spectra suitable for SATRAS analysis are not readily
available.
3.5. MAS and static spectra analysis (MASSA)
SATRAS or even SSTMAS analysis may be completely impossible for amorphous or disordered systems
where lines are often considerably broadened and it is
impossible to obtain the full NMR spectrum with all
the satellite transitions present. In such cases only the
central transition can normally be observed. Facing this
problem Shubin et al. [100] have developed a technique
based on simultaneous analysis of several MAS spectra
recorded with dierent spinning speeds and the static
spectrum. In some experimental situations this approach,
called magic-angle spinning and static spectra analysis, or
MASSA, allows one to obtain the full set of NMR
parameters for quadrupolar nuclei with reasonable
accuracy.
In contrast to analyzing only the intensities of MAS
spinning sidebands as in SATRAS, the MASSA technique
emphasizes the importance of the complete lineshape analysis involving both MAS and static spectra. In this case not

140

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

=
C

B
A

1
2
3
-580

6000

-600

ppm

4000

-620

-640

2000

-600

-2000

-610

-620
ppm

-4000

-6000

-630

-8000

, ppm

Fig. 9. 51V MAS NMR spectra of LaVO4 at 9.4 T, mr = 10 kHz. (A) Full experimental spectrum. (B) The isotropic line (solid) shown with the simulation
(dotted) calculated with the following parameters: CQ = 5 MHz, gQ = 0.4, dr = 200 ppm, gr = 0.7, a = 20, b = 20, c = 0. (C) Simulated isotropic line at
dierent values of dr and b, (1) dr = 172 ppm and b = 88, (2) dr = 272 ppm and b = 88, (3) dr = 272 ppm, b = 8.

only singularities in the line shape are taken into account,


but the complete MAS and static lineshapes are simultaneously computed. Another advantage of the MASSA
technique is its applicability in situations when the CSA
and second-order quadrupolar eects are comparable.
However, the MASSA technique still remains computationally intensive.
An example of MASSA analysis is presented in
Fig. 10 for a VOx/TiO2 catalyst (20 wt% V2O5) [100].
This catalyst was prepared by the spray drying technique
[101]. In such catalysts vanadium is present on the surface in the form of strongly bound vanadium (SBV), as
discussed below in Section 9.1.3. 51V MAS spectra
recorded for this sample are extremely complex
(Fig. 10) and very dicult to analyze using traditional
approaches. Since only the central transition was
observed, neither SATRAS nor SSTMAS techniques
were applicable. Any attempts to represent experimental
MAS spectra recorded at dierent spinning speeds by a
single set of several overlapping resonances also failed.
This suggested that in this case the complex lineshape
of MAS spinning sidebands was determined to a large
extent by the second-order quadrupolar eects. The magnitudes and the relative orientation of 51V quadrupolar
coupling and CSA tensors could be determined only by
tting NMR parameters using simultaneously static and
MAS spectra. A good agreement was observed between
all experimental and simulated spectra tted with a single
set of parameters as illustrated in Fig. 10. This example
shows that inclusion of a static spectrum into the optimization procedure is very desirable.

4
3

2
1
0

-500
-1000
, ppm

-1500

Fig. 10. Experimental and calculated 51V NMR spectra of a VOx/TiO2


catalyst (20 wt% V2O5) at 9.4 T. (1) Simulated static spectrum. (2)
Experimental static spinecho spectrum. (3) Simulated 12.2 kHz MAS
spectrum. (4) Experimental 12.2 kHz MAS spectrum. (5) Simulated
14.1 kHz MAS spectrum. (6) Experimental 14.1 kHz MAS spectrum.
Experimental static and 12.2 kHz MAS spectra, (2 and 4), were simultaneously used in the optimization procedure. The nal simulation of all
three calculated spectra, (1, 3, and 5), was calculated using a single set of
NMR parameters as following: CQ = 14.7 MHz, gQ = 0.59, dr = 650 ppm,
gr = 0.02, diso = 611 ppm, d1 = 281 ppm, d2 = 292 ppm, d3 = 1261
ppm, a = 20, b = 62, c = 42.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

3.6. Multi-quantum MAS (MQMAS) and satellite


transition MAS (STMAS)
The multi-quantum MAS technique is the most suitable
for quadrupolar nuclei with small magnetic shielding
anisotropy and pronounced second-order quadrupolar
eects [102,103], as often observed, for example, for 27Al.
However, for 51V, and for 93Nb in lesser extent, the magnetic shielding anisotropy can be very signicant. In such
cases the MQMAS technique requires more accurate setting up, with carefully chosen pulse lengths and delays.
The most popular 3QMAS pulse sequence is composed
of three pulses instead of two pulses in the originally pro-

p1

p3

t1

posed sequence, where the third p/2 central transition selective pulse is added to the rst two pulses. This pulse
sequence is shown in Fig. 11A from Ref. [104]. The third
selective pulse is used to lter out all satellite transitions
that are being excited by the second pulse, and allowing
through only the central transition, using this so-called
zero-quantum lter.
For example, the optimal excitation of the triple-quantum coherence for spins I = 7/2 and 9/2 nuclei can be
achieved by applying 120 and 90 rf pulses [105]. This is
illustrated in Fig. 12A, where the theoretical and experimental excitation proles are presented depending on the
pulse length for a case of mrf/mQ = 1.25, 2.5, 5. All of them

p4

141

t2

I
+3
2
3

-1

S
n

-3

p1

p3

p4

t1

t2

I
+1

S
rotor
0

p1

p2

-1

-1

p3

t1

p4

t2

I
+2

+1

-1

-1

-2

a
2

x yx yyxx y

xyx yyxxy

2nr

rotor
0

10

Fig. 11. Typical pulse sequences for high-resolution 2D NMR of quadrupolar nuclei. (A) MQ(3Q)MAS, (B) STMAS, (C) DQ STMAS, (D) QCPMG, (E)
SEDOR (low-resolution technique for static samples), (F) REDOR, (G) TRAPDOR, (H) REAPDOR. Radio frequency pulses are dened as p1
excitation, p2 CT-selective, p3 z-lter mixing, and p4 detection pulse. Reproduced with permission from Refs. [81,104,115].

142

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

2.0

7/2

0.3

7/2

1.5

0.2
1.0

0.1
1/2, 1/2

0.0

0.0

3/2, 3/2

0.5

2.0
9/2

9/2

0.3

1.5

0.2
1.0

0.1

0.5

0.0

0.0
0

60

120 180 240 300 360


o
1 ( )

20 40 60 80 100 120 140


o
2 ( )

Fig. 12. Optimization of MQMAS experiments for spin-7/2 and spin-9/2 nuclei. (A) Buildup curves for generation of triple-quantum coherence as a
function of the pulse angle h1. (B) Dependence of triple- to single-quantum coherence transfer processes on the rf pulse angle h2. Investigations focused on
the 0 +3(t1) 1(t2) transfer pathway, and employed initial excitation pulses h1 = 120 for I = 7/2 and h1 = 90 for I = 9/2. Computation assumed the
following ratios between mQ and mrf (gQ = 0): mq/mrf = 5 (--),mQ/mrf = 2.5 (- - -), mQ/mrf = 1.25 (). Experimental data (s) collected for the latter ratio with
93
K59
3 CoCN5 (I = 7/2) and Li NbO3 (I = 9/2) samples at 9.4 T, and subsequently normalized to match the theoretical proles. Reproduced with
permission from Ref. [105].

are normalized to the intensity of the central +1/2 M  1/2


transition, i.e. the maximum intensity of the +3/2 M 3/2
satellite transition is 50% greater that the central transition.
While performing MQMAS experiments, it is important
to achieve the most complete conversion of the triple-quantum coherence, evolving during t1 time, into detected
single-quantum coherence. The nal intensity of the single-quantum coherence as a function of the conversion
pulse is shown in Fig. 12B, with the optimal conversion
pulse found to be 45 for I = 7/2 and 35 for I = 9/2.
Because the intensity of the observed signal is now three
times lower than the intensity of the signal from the central
transition alone, the dierence in sensitivity needs to be
compensated by increasing the number of accumulated
scans by 32 times. The relaxation time of the triple-quantum coherence is proportional to the coherence order, in
this case by a factor of 3. If this is also taken into account
for each increment in t1 evolution time, then 33 = 27 times
the number of scans need to be accumulated in order to
obtain the same signal-to-noise level as in a simple onedimensional single-pulse spectrum. This should be further
multiplied by the number of increments in the t1 direction:
at least 32 or more increments are usually required.
It is clear, that two-dimensional MQMAS spectra
require long recording times. Attempts to cut acquisition
times by decreasing the number of scans would usually
result in a poor signal-to-noise ratio. A simple estimate
shows that when a simple one-dimensional spectrum can
be acquired in as fast as 60 s, MQMAS for the same compound may require several hours. This makes MQMAS
not very feasible for samples with less than 10 wt% or so
of vanadium or niobium content, which is typical for solid
catalysts. All this renders the use of MQMAS very limited
in studying catalytic systems.

However, MQMAS can still be quite useful, particularly


in cases where there are several sites with close isotropic
chemical shifts. As an example, a 51V 3QMAS spectrum
recorded for AlVO4 is shown in Fig. 13A. In this compound there are three non-equivalent vanadium sites.
These three sites are completely resolved in 3QMAS spectra, and the quadrupolar coupling parameters can also be
obtained for each site. Analysis of the one-dimensional
NMR spectrum for this compound is complicated by the
signals being strongly overlapping. The MASSA technique
is not very ecient in this case due to relatively weak second-order quadrupolar eects, and SATRAS is complicated by over 120 individual spinning sidebands needing
to be resolved, further attributed to each particular transition, and nally integrated.
Recently a two-pulse MQMAS sequence has been suggested for measuring CSA parameters when the magnetic
shielding is small and cannot be determined from onedimensional spectra with enough precision [106]. Indeed,
it is very dicult to accurately measure the 51V chemical
shift anisotropy when it is less than 100 ppm, since at this
magnitude the dipolar interactions often conceal any magnetic shielding eects. Taken without any additional
transformations, a 3Q projection of the 3QMAS spectrum
is eectively a spectrum recorded at three times the spectrometers magnetic eld. Because the CSA interactions
measured in parts per million do not depend on the magnetic eld strength while the dipolar interactions are
inversely proportional to the magnetic eld on the parts
per million scale, the number of spinning sidebands in
the 3Q dimension is increased three times at the same
spinning speed. This improves accuracy in determining
the chemical shift anisotropy as well as the asymmetry
parameter.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

id 51V, ppm
-800

143

id 93Nb , ppm
-1600

V3

A
V2

-1200

-750

Nb 1
-700

-800

V1
-400

-650
-650

-700

-750

-800

-900

51V, ppm
id

51

id

V, ppm

-640

-1000

93

-1100

-1200

Nb, ppm

Nb , ppm

-1050

93

CT

-620

Nb 1

-600

-1000

-580

ST1
-560

ST2

-950

-540
-590

-600

-610

51V, ppm

-620

-1150

-1200

-1250

-1300

93Nb, ppm

Fig. 13. (A) 51V 3QMAS sheared spectrum of AlVO4 obtained at 9.4 T. (B) 93Nb 3QMAS sheared spectrum of BiNbO4 obtained at 9.4 T. (C) 51V STMAS
sheared spectrum of LaVO4 showing dierent correlations. CT is the autocorrelation of the central transition, ST1 correlates the central transition with the
1/2 M 3/2 satellite transition, and ST2 correlates the central transition with the 3/2 M 5/2 satellite transition. The experiment has been recorded at
9.4 T using 10 kHz MAS. (D) 93Nb DQ STMAS spectrum of Te3Nb2O11 recorded at 21.1 T using 20 kHz MAS.
93

Nb MQMAS spectra are usually not complicated by


CSA. At the same time, the second-order quadrupolar
eects are very pronounced, thus simplifying the experimental setup compared with 51V [50]. Even at lower spinning speeds the spectral resolution in MQMAS spectra
can sometimes be improved because of using the second
dimension that spreads out overlapping resonances and
MAS spinning sidebands. An example of 93Nb 3QMAS
NMR is shown in Fig. 13B for BiNbO4.
To summarize, the main advantage of the MQMAS
technique, although being inherently of low sensitivity
and time-consuming, is increased spectral resolution, and
oering the possibility to determine the isotropic chemical
shifts and the quadrupolar coupling parameters in a
straightforward simplied manner.
The satellite transition MAS (STMAS) [107,108] technique has been proposed as a method complementary to
MQMAS. STMAS provides better sensitivity and is less
dependent on CSA and the strength of quadrupolar interactions. As a result the sites with low or high quadrupolar
coupling constants should be easier to observe in STMAS
than in MQMAS. The STMAS has certain similarities with
both MQMAS and SATRAS experiments as it correlates
in a two-dimensional spectrum the central transition with

the satellite transitions (m, m  1). In fact the STMAS


experiment can be described as a two-dimensional
SATRAS experiment. The STMAS experiment requires a
very accurate setting of the magic angle which is also similar to SATRAS [108]. A typical STMAS pulse sequence is
shown in Fig. 11B.
The rst 51V STMAS experiment of LaVO4 was published in [45]. Three signals were observed, corresponding
to the central transition signal (CT), the 1/2 M 3/2
satellite transition signal (ST1), and the 3/2 M 5/2 satellite transition signal (ST2) (Fig. 13C). The composite quad1=2
rupolar coupling constant, k C Q 1 g2Q =3 , and the
isotropic chemical shift, diso, extracted from the position
of the center of gravity of the ST1 line (k = 6.2 MHz,
diso = 605 ppm), and those obtained from the frequencies
of the center of gravity of ST2 (k = 6.0 MHz, diso =
600 ppm) agreed well with SATRAS and MQMAS data
[45]. However, this multiplicity of each signal may complicate interpretation of the spectra, especially when several
sites are involved.
Fig. 11C shows a double-quantum STMAS sequence,
DQ STMAS. A simple modication of the standard
STMAS sequence by placing a selective p pulse before
the t1 period correlates the double-quantum satellite transi-

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

tions with the central transitions, DQ1 CT. This correlation also refocuses the anisotropic second-order quadrupolar interactions yielding ridge-shaped resonances in the
2D spectra. The double-quantum STMAS spectra are free
of diagonal and outer satellite transition peaks similar as in
ltered single-quantum spectra.
The rotor synchronization is important for ecient ltering of the diagonal and outer satellite transition peaks
in DQ STMAS spectra [66,67,104]. The undesired signals
from CT and ST2 can leak through the double-quantum
lter because the soft p pulse (p2, p4 in Fig. 11C) can induce
some coherence transfer to double-quantum despite their
low eciencies. The central transition has a zero rst-order
quadrupolar shift and the outer satellite transitions ST2
have a rst-order shift twice that of ST1. Therefore these
coherences are not refocused at the mixing pulse if the t1
evolution time is carefully chosen for the rotor-synchronization of the rst-satellite transition coherence. The combined eects from low coherence transfer eciencies and
rotor-de-synchronization of these leak-through coherences
explain the superb performance on the suppression of the
CT and ST2 peaks in DQ STMAS, as illustrated in
Fig. 13D for Te3Nb2O11.
3.7. Heteronuclear correlation spectroscopy (HETCOR)
HETCOR spectroscopy is a classic two-dimensional
NMR correlation technique developed to probe shortrange correlations and distances between heteronuclei,
and is very popular in liquid-state NMR spectroscopy.
The recent development of the necessary hardware has
made HETCOR also useful in solid-state NMR [109
112], including research in vanadia and niobia based catalysts. Potentially HETCOR may become an important tool
in the solid-state NMR practice.
The most popular HETCOR approach in solid state is
based on cross-polarization (CP). Even though one-dimensional CP/MAS is not always successful in measuring correlations between protons and quadrupolar nuclei due to very
fast relaxation of the latter, HETCOR allows these correlations to be measured, as for example, for 51V1H pairs.
For spin one-half nuclei CP matching conditions are
described by the HartmannHahn equation, cHB1H =
cXB1X, [113], where B1H and B1X are the rf elds of the
two nuclei during the contact time. A more general HartmannHahn rule applicable to quadrupolar nuclei requires
the nutation frequency of the quadrupolar nucleus to
match the eective proton rf eld: cHB1H = xnut. The
quadrupolar nutation frequency, xnut, depends on many
parameters, including the quadrupolar coupling constant
and the asymmetry of the quadrupolar coupling tensor.
All this complicates dramatically the optimization of CP
matching conditions, which becomes even worse when
there are several non-equivalent quadrupolar sites with different nutation frequencies in the system.
Another problem while using HETCOR arises due to
very short relaxation times caused by the quadrupolar

relaxation mechanism, which result in mixing signals from


nuclei connected to protons and nuclei not connected to
protons. This complication is somewhat circumvented by
using complex cycling pulse sequences, which are often
rotor-synchronized.
An example of a 51V1H HETCOR experiment is shown
in Fig. 14 for Ba(VO3)2H2O. According to the crystal
structure, there are two non-equivalent vanadium sites in
this compound with similar tetragonal pyramid oxygen
environments. However, vanadiumproton distances for
. This
the two sites are quite dierent, 2.005 and 2.727 A
1
51
may explain why the eciency of the H to V magnetization transfer is also dierent for two sites. A simple
one-dimensional 51V MAS NMR spectrum of Ba(VO3)2
H2O has two lines corresponding to two vanadium sites.
These one-dimensional spectra were used to optimize the
HETCOR experimental conditions for this compound
(Fig. 14).
It is also interesting to note, that although the onedimensional 1H MAS NMR spectrum has only a single
peak, the HETCOR spectrum contains a second component of lower intensity from HV interactions corresponding to the second vanadium site. These signals were simply
overlapping in the 1H MAS NMR spectrum.
One of the advantages of the 51V1H HETCOR NMR
technique is the possibility of studying VH bonding, as
well as increased spectral resolution due to the second
dimension. As for many other two-dimensional techniques,
HETCOR suers from low sensitivity, thus requiring
longer acquisition times.
3.8. Double resonance experiments (SEDOR, REDOR,
TRAPDOR, and REAPDOR)
Kaplan and Hahn [114] proposed a spinecho double
resonance (SEDOR) pulse sequence to probe connectivity
and interatomic distances between adjacent nuclei in solids.
A xed-time variation of the SEDOR pulse sequence is
shown in Fig. 11E [115]. SEDOR experiments are
performed on stationary samples and as such have only

-5
1
( H), ppm

144

-600

- 620

- 640

(51V), ppm

Fig. 14. 51V1H HETCOR NMR spectrum of Ba(VO3)2H2O obtained


using 7 kHz MAS at 9.4 T.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

limited use in 51V NMR spectroscopy of solids due to lack


of spectral resolution.
Follstaedt and Slichter [116] applied 51V63Cu SEDOR
to identify 51V impurities dissolved in metallic copper.
They were able to resolve two 51V resonances in CuV solid
solutions belonging to isolated impurities in copper and
undissolved vanadium metal remaining in samples.
SEDOR data were also useful for identication host 63Cu
nuclei adjacent to the vanadium impurity.
The rotational echo double resonance (REDOR) experiments are similar to SEDOR, except that the spin I echoes
are now synchronized with magic-angle spinning (Fig. 11F)
[117]. Brown et al. [118] have used the REDOR technique
for a 51V/15N pair to study phenylamide groups bonded to
VOx species on a silica surface. REDOR has allowed the
determination of a 51V15N dipolar coupling constant
and hence the bond length. The eciency of the polarization transfer was increased by additional 15N/1H cross
polarization. These and similar REDOR experiments
require triple-channel NMR probes.

145

Other examples of REDOR involving 51V have been


reported by Hudalla et al. [119]. They studied 31P{51V}
and 51V{31P} REDOR spectra of ZrV2xPx O7 solid solutions applying two-pulse sequences with two and four rotation cycles. The authors determined the VP bond length in
.
this system to be 3.42 A
TRAPDOR (transfer of populations in double resonance), another variety of the double resonance NMR
technique, has been proposed by Grey et al. [120,121].
TRAPDOR involves nuclear polarization transfer similar
to CP/MAS, however, the transfer is now performed under
adiabatic conditions (Fig. 11G). Kim and Grey [122] used
51
V17O TRAPDOR NMR to conrm the assignment of
resonances in the disordered anionic conductors aBi4V2O11 and c-Bi4V1.7Ti0.3O10.85. TRAPDOR has also
been shown to be sensitive to probe mobility involving
oxide ion hops between equivalent crystallographic sites.
The pulse sequence for a REAPDOR (rotational echo
adiabatic passage double resonance) experiment is shown
in Fig. 11H [123]. This experiment combines REDOR

Fig. 15. 29Si CP/MAS spectrum of [AsW9O33(t-BuSiO)3(VO)]3 and the corresponding ORTEP view (orange, V; gray, As; blue, W; red, O; black, C;
green, Si). Dashed inset shows isotropic lines without {51V} decoupling. Full line inset shows isotropic lines with CW {51V} decoupling. Spinning
sidebands are marked with asterisk. Reproduced with permission from Ref. [125].

146

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

and TRAPDOR approaches, but is less susceptible to


errors when conditions of a perfect adiabatic passage are
not met. Huang et al. [124] have demonstrated that the
REAPDOR spectroscopy is applicable to distance
measurements between spin-1/2 and -7/2 pairs, i.e. 31P 51V.
Using REAPDOR they measured PV distances in monovanadium substituted K4PVW11O40, 1-K7P2VW17O62, and
4-K7P2VW17O62 polyoxoanionic salts. Numerical simulations of the experimental NMR data yielded very good
agreement with the averaged PW/PV distances determined from the X-ray diraction measurements in the same
or related compounds.
3.9. Triple Resonance experiments
Only a limited number of triple resonance experiments
involving 51V or 93Nb have been reported so far. A series
of 51V/29Si/1H, 31P/13C/1H, and 29Si/13C/1H experiments
were performed by Bonhomme et al. [125] to study a
[AsW9O33(t-BuSiO)3(VO)](n-Bu4N)3 complex. In their
experiment they detected 29Si magnetization decay, while
the population transfer was performed from 1H via 51V.
This approach resulted in dramatically improved spectral
resolution, allowing direct observation of 51V29Si splitting
in the 29Si spectra from a scalar coupling with the value
J(51V29Si) = 28 Hz (Fig. 15). We note here that the twobond heteronuclear J-coupling in 93Nb29Si was rst
observed by Kao and co-authors [126] using simple onepulse experiments for 93Nb and 29Si. They reported
J(93Nb29Si) = 64 Hz in Rb4(NbO)2(Si8O21).
3.10. Advantages of high magnetic eld strengths
The availability of NMR instruments operating at magnetic elds exceeding 1820 T is increasingly attractive for
solid-state NMR research. Besides providing improved
sensitivity, there are many other advantages of using ultrahigh magnetic elds for studying half-integer-spin quadrupolar nuclei [127,128] such as 51V, 93Nb and 181Ta.
Considering quadrupolar coupling interactions, under
stationary conditions the full width of the central transition

-100

line shape, DmCT, broadened due to second-order quadrupolar interactions is inversely proportional to the resonance frequency, m0, as
DmCT Hz / C 2Q =m0

Performing experiments at higher magnetic elds results


in proportionally narrower static line widths in cases
where this type of broadening is dominant. For example,
the full central transition linewidth in the 93Nb spectrum
recorded for a stationary La3NbO7 sample at 9.4 T
(97.8 MHz) is about 225 kHz. When the same spectrum
is recorded at 21.1 T (220.0 MHz) this line is now only
about 100 kHz wide (Fig. 16A), i.e. the scaling factor
is 2.25, very close to what is predicted from the corresponding magnetic eld values. If the two spectra are
compared on the parts per million scale, the narrowing
at higher eld becomes even more dramatic, since on
the parts per million scale the second-order quadrupolar
linewidth of the central transition is inversely proportional to the square of the resonance frequency, i.e. the
scaling factor in this case is 5.1 (Fig. 16B).
This type of narrowing at high magnetic elds simplies
considerably the recording of the static spectra of central
transitions aected by the second-order perturbations.
Narrower static lines at higher magnetic elds oer better
sensitivity and require much less rf power for homogeneous
excitation. Spectra can be obtained much faster than at
lower elds, and in fewer steps if acquired using a
stepped-frequency acquisition technique. Even the largest
quadrupolar coupling constants, earlier accessible only
via nuclear quadrupolar resonance NQR, can now be
directly measured with solid-state NMR.
We note, however, that in contrast to the second-order
broadening, the rst-order quadrupolar eects do not scale
with the resonance frequency, i.e. the full span of the satellite transitions will still cover the same spectral width measured in Hertz regardless of the magnetic eld strength.
As in stationary samples, under MAS conditions the
linewidth of the central transition is also inversely proportional to the resonance frequency. This is illustrated in
Fig. 17 where two 93Nb MAS NMR spectra are compared

-200

kHz

-300

-400

2
1000

-1000

-2000

-3000

, ppm

Fig. 16. Eects of the magnetic eld strength on the 93Nb NMR spectra recorded at 9.4 T (above) and 21.1 T (below) for a stationary La3NbO7 sample.
(A) Spectra scaled in kilohertz. (B) Spectra scaled in parts per million.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

147

21.1 T

50

-50

-100

-150

-200

-250

9.4 T

-300

kHz

Fig. 17. Eects of the magnetic eld strength on 30 kHz MAS 93Nb NMR
spectra of Te3Nb2O11 recorded at 9.4 T (below) and 21.1 T (above).

for Te3Nb2O11. Even though both spectra were recorded


with the same MAS spinning speed (30 kHz), the spectrum
obtained at 21.1 T is considerably simplied comparing
with the spectrum recorded at 9.4 T.
This narrowing at higher magnetic elds makes MAS
experiments feasible in systems where the quadrupolar coupling constants are quite large. Improved spectral resolution in MAS spectra recorded at high elds becomes of
great importance when there are several sites present,
which is very typical for complex multi-component systems, including many advanced materials and catalysts,
as will be discussed in more details below. Here we illustrate this with 51V MAS NMR spectra obtained for a
2.5% V2O5/ZrO2 sample at two magnetic elds, 9.4 and
21.1 T (Fig. 18).
Even though, the spectrum at 9.4 T was recorded when
spinning the sample at 35 kHz the resolution in this spectrum is much lower than in the 18 kHz MAS spectrum
obtained at 21.1 T. In the latter spectrum up to ve individual sub-spectra can be resolved from ve dierent types of
VOx species on the ZrO2 surface. It is clear that the NMR
spectra recorded for these and similar systems at high and
ultrahigh magnetic elds contain more useful structural
information, which can be used to improve their properties
and performance.
Important structural information can be obtained from
studying magnetic shielding tensors as well. This is also
facilitated at ultrahigh magnetic elds, since the chemical
shielding anisotropy measured in Hertz scales linearly with
the magnetic eld strength. Very small chemical shielding
anisotropies become more apparent at higher magnetic
elds, even if at lower elds they are eectively concealed
by other types of line broadening, including quadrupolar
or dipolar interactions. For example in a static 51V NMR
spectrum recorded for Ba3(VO4)2 at 9.4 T the central transition line is rather broad and featureless mostly due to

-500

-1000

-1500

, ppm

Fig. 18. 51V NMR spectra of a 2.5% V2O5/ZrO2 catalysts. (1) Static
spectrum at 9.4 T. (2) Simulated spectrum (1). (3) 35 kHz MAS spectrum at
9.4 T. (4) 18 kHz MAS spectrum at 21.1 T. (5) Simulated spectrum of (4).

-580

-600

-620

-640

-660

-580

-600

-620

-640

-660

-200

-400

-600
-800
, ppm

-1000

-1200

Fig. 19. 51V NMR static spectra of Ba3(VO4)2. (1) Recorded at 9.4 T.
Inset shows the central transition. (2) Recorded at 21.1 T. Inset shows the
central transition (upper spectrum) together with its simulation (lower
spectrum).

quadrupolar broadening (Fig. 19). These broadening


eects diminish at higher magnetic elds, and the same
spectrum recorded at 21.1 T shows now the central transition line typical for an axial anisotropy of the chemical
shielding tensor, which can be simulated, and the anisotropy parameters extracted.
The fact that the chemical shielding anisotropy scales
linearly with the magnetic eld is particularly important
for 93Nb NMR, where even at reasonably high elds of
714 T the static spectra are often still dominated by the

148

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

quadrupolar interactions. Only at ultrahigh magnetic elds


do the CSA eects become clearly visible in such spectra
and can accurately be measured as will be demonstrated
below. Ultrahigh eld NMR spectroscopy is gaining in
popularity for studying CSA eects in a variety of quadrupolar nuclei [129131].
On a cautionary note, the very same fact that the chemical shielding anisotropy measured in Hertz scales linearly
with the magnetic eld may complicate recording MAS
spectra at very high magnetic elds. This is particularly
important to remember while performing 51V MAS
NMR experiments, where spectra are often dominated by
the magnetic shielding interactions. In such cases to achieve
the same resolution on the parts per million scale in MAS
spectra it is often necessary at higher magnetic elds to spin
samples at much faster MAS spinning speeds. For example, to just reproduce on the parts per million scale a spinning sideband pattern originating from CSA in the 35 kHz
MAS spectrum recorded at 9.4 T, it would require spinning
the same sample at 79 kHz at 21.1 T. This is not always
possible in practice.
Another complication arises when the line width is dominated by a chemical shift distribution. This situation is typical for many glasses and amorphous materials. In such cases
employing ultrahigh magnetic elds does not necessarily
result in improved spectral resolution. Because the chemical
shift distribution is not aected by the magnetic eld strength
or the resonance frequency, the line width measured in parts
per million remains the same regardless of the eld. Moreover, when measured in Hertz, the line dominated by the
chemical shift distribution will actually be broader at higher
eld. However, the spectral resolution may still improve with
increasing eld if T2 values do not vary strongly as a function
of applied magnetic eld strength.
Two other types of interactions to be mentioned here,
dipolar and scalar J-coupling, are eld-independent when
measured in Hertz, and in solid-state 51V and 93Nb NMR
should be studied at moderate magnetic elds. At ultrahigh
elds these two interactions are often concealed by broader
chemical shift distributions, particularly in materials which
are not of extremely high crystallinity.
Paramagnetic eects in 51V and 93Nb NMR spectra will
be discussed in more details below. Here we mention only
that the frequency contributions to the magnetic shielding
from both the Fermi contact and pseudocontact interactions are directly proportional to the magnetic eld
strength. At high magnetic elds this may cause severe line
broadening and even disappearance of some resonances
aected by paramagnetic interactions. In paramagnetic systems employing ultrahigh magnetic elds does not necessarily lead to better or improved spectra quality.
4. DFT and other quantum chemical computational
approaches
Theoretical computations of NMR parameters (CSA
and EFG tensors) in transition metal compounds represent

a separate topic of quantum chemical research. We refer


the reader to textbooks and introductory articles [132
134] and to a recent comprehensive review [135]. Chemical
shift and EFG computations by the density functional theory (DFT) approach are implemented in some popular
quantum-chemical packages. DFT methods are now in
the most commonly used for the theoretical prediction of
NMR parameters for nuclei with high atomic number it
was concluded by Autschbach [135] that the HartreeFock
approach is, in general, not adequate, except maybe for the
special case of high metal oxidation states of d10 systems.
We also note that relativistic eects can have a serious
inuence on calculated NMR parameters. Non-relativistic
calculations of the chemical shift in metals usually produce
reliable results only for 3d and 4d series elements, while for
5d and f-block metals a proper treatment of relativistic
eects is mandatory. Even for 4d-block elements relativistic
corrections to the absolute shieldings can be signicant, but
they usually have similar value for dierent computed
structures and, therefore, are almost canceled in relative
chemical shift calculation being masked by the broader
range of the relative chemical shifts. Even though just
few in number, all recent theoretical predictions of 51V
NMR properties in solid state have also been performed
with the use of the DFT method.
Using two dierent basis sets Pooransingh et al.
[136,137] and Ooms et al. [138] computed the quadrupolar
and chemical shielding anisotropies by a DFT method for
four crystallographically characterized oxovanadium (V)
complexes mimicking the active site of vanadium haloperoxidases and hydroxylamido. The calculated tensors are in
general agreement with the experimental solid-state NMR
data. It was concluded that the combination of 51V solidstate NMR and computational methods is benecial for
investigating electrostatic and geometric environments in
diamagnetic vanadium systems with moderate quadrupolar
anisotropies.
Based on isolated and embedded cluster models Gee
reported [139] DFT calculations of the 51V EFG tensor elements in orthovanadates to be in very good agreement with
experimental results. It was concluded that the local vanadiumoxygen covalent bond structure had a more pronounced eect on the 51V EFG than did the rest of the
crystal. This is in agreement with earlier DFT calculations
based on the linearized augmented plane-wave method for
periodic solids [140]. In this latter work, DFT calculations
provided reliable assignment of 27Al and 51V resonances to
specic crystallographic sites in the asymmetric unit of
AlVO4. It was shown that the magnitude and orientation
of the EFG tensors are largely determined by the
p  p(27Al) and p  p, d  d(51V) orbital contributions to
the valence electrons, while the lattice itself provides only
minor contributions for both nuclei.
Despite continuous development of quantum chemical
methods, reliable calculations of shielding parameters for
transition metals have not been possible until recently.
The approaches commonly used before were based on

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

nite, in a number of atoms, models using localized


atomic orbitals. In a recently published study [141] periodic DFT calculations of 51V NMR shielding parameters
were reported for the rst time employing the gaugeincluding projected augmented-wave (GIPAW) pseudopotential approach [142]. Using pseudopotentials specically constructed for vanadium Truandier et al. [141]
were able to perform accurate periodic calculations of
the 51V shielding tensor in AlVO4 as well as to perform
EFG calculations. Analysis of the relative orientation of
the EFG and shielding tensors has shown that these
rst-principles calculations can indeed give access to
Euler angles describing relative orientation of these two
tensors.
Most recently an ultrasoft pseudopotential (USPP)
modication of the GIPAW method has been applied to
systematically calculate 49Ti and 51V shielding parameters
in a series of molecular and extended periodic systems
[143]. Thirteen dierent vanadate compounds, with total
of 18 distinct vanadium sites, were computed, and a very
good correlation was found between theoretical and experimental values of isotropic chemical shift. Obtained data
allowed straightforward assignment of 51V resonances in
AlVO4, a- and b-Mg2V2O7, and Ca2V2O7 compounds.
Quite reasonable agreement between experimental and theoretical CS anisotropy parameters was also obtained for
ortho- and pyrovanadates.

Table 4
Dierent types of vanadium sites

5.

51

149

V NMR data compilation

5.1. Chemical shielding and quadrupolar tensor parameters


in individual vanadium compounds
The vanadium coordination number in vanadium-containing oxide compounds varies from 4 to 6 with vanadium
found in dierent VOx environments. For example, VO4
sites (x = 4) may exist either as tetrahedral sites (Td) or trigonal pyramids (Pd) (Table 4). Association of VOx polyhedra is often described as Qn, where n corresponds to the
number of shared oxygen atoms. Thus tetrahedral Q0 sites
(n = 0) are isolated species, while Q1 are dimers (n = 1),
and Q2 (n = 2) correspond to chains of VO4. A similar situation exists for VO5 and VO6 species. However, the level
of association of VO5 and VO6 polyhedra can be higher.
For example, VO6 are isolated at n = 0, dimeric at n = 1
and 2, form chains at n = 2, 3, 4, layers at n = 4, and polyoxoanions at n = 6 (Table 4).
Precisely measured 51V NMR parameters are currently
available for many individual binary vanadium oxide compounds of compositions V2O5MxOy, where M is mono-,
di-, tri-, tetra- or pentavalent metal, or V2O5XzOk, where
X is phosphorus, arsenic or sulfur. Some of the ternary
vanadium compounds, i.e. V2 O5MxOyXzOk, and even
several V-containing biological systems, have also recently
been studied [93,94,144149]. It is noteworthy, that almost

150

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

all types of vanadium sites shown in Table 4 can be found


in these individual compounds.
5.1.1. Tetrahedral Q0 sites
Sites Q0 correspond to vanadium in tetrahedral oxygen
coordination where individual VO4 3 tetrahedra are isolated from each other. This type of vanadium coordination
is typical for almost all known orthovanadates:
M+: Li3VO4 [150], Na3VO4, K3VO4 [151], Cs3VO4
[152,153]; Tl3VO4.
M2+: Ca3(VO4)2 [154], Sr3(VO4)2 [155], Ba3(VO4)2
[156], Mg3(VO4)2, Zn3(VO4)2, Pb3(VO4)2 [157].
M3+: AlVO4 [158], BiVO4, LaVO4 [159,160], LuVO4,
YVO4 [161], ScVO4 [161], CeVO4 [162], RhVO4.
M5+: TaVO5, NbVO5 [163], VNb9O25, VTa9O25 [164].
In many of these compounds VO4 3 tetrahedra are
almost ideally symmetric with VO distances in the range
, and all these distances usually within less then
of 1.61.8 A

0.1 A of the average VO distance.


As an example, experimental and calculated 51V MAS
NMR spectra of Li3VO4 are presented in Fig. 20. The

3
4

5
2000

1000

, ppm

-1000

-2000

-3000

Fig. 20. 51V MAS NMR spectra of Li3VO4 at 9.4 T. (1) Experimental
spectrum. (2) Simulated spectrum with parameters: m0 = 105.2 MHz,
CQ = 1.53 MHz, gQ = 0.05, dr = 12 ppm, gr = 0.8, diso = 544 ppm.
(35) Simulated satellite transitions 7/2 M 5/2, 5/2 M 3/2, and
3/2 M 1/2 correspondingly.

experimental spectrum obtained for this compound under


spinning at 12.5 kHz has very high resolution, sucient
to detect and identify a minor, less than 1%, impurity of
LiVO3. The 51V MAS NMR spectrum of Li3VO4 represents a well-resolved superposition of six individual subspectra from six satellite transitions, 5/2 M 7/2,
3/2 M 5/2, and 1/2 M 3/2. These spectra can be
analyzed within the SATRAS approach as described
above. CSA and quadrupolar parameters determined for
this compound from SATRAS analysis are summarized
in Table 5. The value of the quadrupolar coupling constant
found for this compound is rather small, CQ = 1.53 MHz,
which explains why the second-order quadrupolar eects
are virtually undetectable in this case.
Similar spectra have been obtained and analyzed for all
studied orthovanadates. Based on observed 51V NMR
parameters (Table 5) it is possible to conclude that vanadium in regular Q0 tetrahedral oxygen environment has
almost spherically symmetric magnetic shielding tensor
with small CSA anisotropy, dr < 100 ppm. The quadrupolar coupling constant, CQ, is found in the range of 1
6 MHz, while the asymmetry parameter of the quadrupolar
tensor varies from 0 to 1 depending on the compound
[93,94,149].
5.1.2. Tetrahedral Q1 sites
Vanadium in tetrahedral oxygen coordination of Q1
type is sharing an oxygen atom with adjacent VO4 3 and
forming a V2 O7 4 dimer with two VO3 units connected
via a bridging VOl2V link. Sites Q1 are typical in pyrovanadates of monovalent and divalent metals. Q1 tetrahedra in pyrovanadates are usually somewhat distorted
compared with more symmetric Q0 units in orthovanadates. A slight distortion in Q1 is caused by the bridging
VOl2 bond, which is normally longer than the three
remaining VOt bonds.
In pyrovanadates of divalent metals two vanadium
atoms in V2 O7 4 are usually non-equivalent. Depending
on the V2 O7 4 structure two classes of pyrovanadates can
be distinguished. In pyrovanadates of the thortveitite type
(structural analogs of Sc2Si2O7) the VOl2V angle is at
180, i.e. the V2 O7 4 dimer has D3d symmetry, as for example, is observed in a-Zn2V2O7 and Cd2V2O7. At the same
time in pyrovanadates of the dichromate type (structural
analogs of K2Cr2O7) the VOl2V angle is at about 140
and the V2 O7 4 dimer has C2v symmetry, i.e. as found in
a-Mg2V2O7, b-Mg2V2O7, Ca2V2O7, BaCaV2O7, and aBaZnV2O7. 51V NMR parameters for many pyrovanadates
of divalent metals have been reported by Nielsen et al.
[147].
For two thortveitite compounds, a-Zn2V2O7 and
Cd2V2O7, the values of CQ were determined as 3.9 and
6.0 MHz, and the value of dr were found to be 119 and
173 ppm. For dichromate compounds CQ is found within
a wider range of values from 1.5 to 10 MHz, while there are
two general regions for dr, one is from 262 to 57 ppm,
and the second is from 70 to 250 ppm.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

151

Table 5
51
V NMR parameters of individual vanadium compounds
Compound
VO4 type Q0
Li3VO4

Na3VO4

Ref.

Site

CQ
(MHz)

[34]
[22]
[149]

1.52
1.51
1.53

[185]
[22]

1.05

gQ

dr
(ppm)
50

0.05

12

[34]

100

Cs3VO4

[34]
[149]

56
30

[34]

Mg3(VO4)2

[145]
[34]
[186]

Ca3(VO4)2

[34]
[22]

Sr3(VO4)2

Ba3(VO4)2

diso
(ppm)

d11
(ppm)

d22
(ppm)

d33
(ppm)

533

543

556

520
555

580
561

626
603

533

550

588

542

557

572

0.80

3.50

0.29

1.05

0.56

c/o

544.3

31
40
15

576
572.9

64

85

61

67

114
137
130

29
27
25

88
58
15

165

77

36

90

30

66

169

480
0.53
1

557.3
557
557
615

2.05

0.53
0.53

0.01

20
14

0.86

610
608(618)

631

619

604

10

0.50

610.9

603

608

620

20

605

0.75
0.76

0.01

11
5

0.94
0.27

603(604)
602.9

593
600

604
601

616
608

0.99

41

0.68

488

516

564

Zn3(VO4)2

[145]
[185]
[186]

1.04

28

0.93

522.5
522
522(520)

Pb3(VO4)2

[186]

1.01a

493

519

548

32

0.41

486(498)

521

508

467

Mg2Sr(VO4)2
Mg2Ba(VO4)2

[186]

38

0.66

581(586)

617

592

548

[186]

37

0.16

575(584)

606

600

547

Mg2Pb(VO4)2

[186]

51

0.18

549(544)

574

565

493

AlVO4

[34]

62
55
75
90
90
90
87
120
82

0.16
1.00
0.40
0.30
0.30
0.30
0.74
0.72
0.88

668
747
780
662.0
743.0
776.0
660.5
743.6
775.7

630
605
710
604
685
718
585
847
853

640
745
800
631
712
745
649
760
781

730
800
830
752
833
866
748
624
694

40

0.83

674.9

638

672

715

665.8

644

648

706

609

555

616

657

604.7
612.1

562
573

597
579

655
684

1
2
3
1
2
3
1
2
3

3.50
3.30
4.50
4.05
2.35
3.08

0.70
0.70
0.70
0.84
0.93
0.62

3.95

0.01

ScVO4

[149]

YVO4

[34]
[22]
[149]

4.75
4.85

0
0.01

[34]
[22]
[106]
[149]

5.21
5.20
5.21

0.69
0.68
0.69

50
72

CeVO4

[188]

5.62

0.21

177

0.01

427

338

339

604

LuVO4

[149]

4.28

0.01

0.43

663.4

660

662

668

BiVO4

[34]
[145]
[189]
[190]

4.94
4.82
5.04

0.36
0.41
0.4

80
94
138
97

0.63
0.32
0.099
0.7

420
421.1
416.8
428

355
359
341
345

405
389
354
414

500
515
555
525

LaVO4

b/o

560
0.82
0.18

100

[34]
[22]
[186]
[149]

[187]

a/o

544

30

[34]
[186]
[22]
[149]

[149]

d^
(ppm)

545

K3VO4

Tl3VO4

gr

30
40
72

664
0
0.10
0.85
0.69
0.71
0.09

RhVO4

[191]

LiCdVO4

[192]

3a

125

NbVO5

[145]

1.2

0.39

70

0.17

791.4

750

762

861

90

25

90

TaVO5

[145]

0.85

0.28

53

0.24

773.2

740

753

826

90

37

90

VNb9O25
VTa9O25

[193]
[193]

3.95
4.5

0.12
0.1

100
160

0.01
0.06

602.0
617

552
532

553
542

702
777

605
630

(continued on next page)

152

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

Table 5 (continued)
Compound

Ref.

Site

VO4 type Q1
Na4V2O7

[34]

1
2

K4V2O7

[34]

Cs4V2O7

[34]

Tl4V2O7

[34]

a-Mg2V2O7

[34]
[147]

b-Mg2V2O7

[34]
[147]
[186]

Ca2V2O7

[34]
[147]

CQ
(MHz)

gQ

[34]

[149]
[186]
Ba2V2O7

[34]

[149]

[186]
a-Zn2V2O7

Cd2V2O7

64

1
2
1
2
1
2
1
2
1
2
1
2
1
2

BaCaV2O7

[147]

ZrV2O7

[34]
[149]

d22
(ppm)

d33
(ppm)

500

582

642

d^
(ppm)

a/o

b/o

c/o

578

80
2

1
89

0
41

0
36

25
4

0
17

8
28

90
90

16
69

71
128

10
0

77
90

131

90

82
95

38
27

90
90

2
132

36
17

90
50

543
567

3.29
4.82

10.10
4.80

1.58
7.33

0.69
0.43

0.44
0.39

0.90
0.43

2.50
3.20

0.85
0.90

2.04
2.37
2.09

0.95
0.85
0.76

3.97
3.86

0.54
0.56

6.00

0.41

52

0.85

504

443

512

556

30
83
57
103

0.50
0.12
0.91
0.34

555
617
549.2
603.5

510
570
604
534

570
580
552
570

585
700
492
707

120
50
262
113
36
73

0.83
0.80
0.10
0.90
0.8
0.68

560
650
494.4
639.3
551(574)
647(655)

460
590
639
747
611
717

560
660
612
645
571
667

680
700
232
526
542
582

56
62
71
530
62

0.64
0.55
0.54
0.50
0.68

574
578
574.9
534.0
575

528
530
520
137
523

564
564
559
402
565

630
640
646
1064
637

43
68
64
66
78

0.58
0.44
0.31
0.30
0.40

523
480
480
480
527

548
620
632
638
559

600
650
652
658
660

83

0.12

557
582
588
592
582.0
561.5
561,581 584,593

542

552

672

73
27
40
51
10
53

0.34
0.04
0.38
0.53
0.95
0.94

579
588
600
596.0
587.0
577.0
579,589,599

530
574
535
635
577
526

555
575
625
608
587
575

652
615
640
545
597
630

120
119
119
117

0.67
0.69
0.62
0.65

620
615
616.6
620(612)

500
514
713
494

640
597
639
632

720
734
498
709

193
173

0.00
0.27

563
562.7

370
673

660
626

660
390

110
89

0.45
0.22

522(510)
521(516)

430
462

480
482

620
606

100
99

0.65
0.49

581.6
598.6

499
525

564
574

682
698

2.443

0.169

110
75

0.05

774
775.9

710
737

802
740

824
851

3.81
5.91

0.58
0.86

143
252

0.10
0.16

650.1
608.4

571
462

586
503

793
860

244

0.34

365

202

285

609

[34]
[93]
[23]
[24]
[25]
[144]

2.98
2.91
2.88
2.95
2.76
2.95

0.28
0.3
0.3
0.19
0.37
0.3

234
237
250

0.64
0.71
0.28

572
563.7

380
361

530
529

807
801

75

23

34

240

0.7

569.5

366

534

810

78

23

32

[144]
[145]

3.18
3.25

0.87
0.88

221
228

0.7
0.72

573.4
573.2

386
377

540
541

794
801

154
105

9
21

4
72

VO4 type Q2
a-AgVO3

[194]

1
2

d11
(ppm)

0.32
0.85

[147]

LiVO3

1
2
3
1
2
3
1,2,3

diso
(ppm)

2.57
3.20

a-BaZnV2O7

NH4VO3

1
2
3
4
1
2,3,4
1,2,3,4

[185]
[147]
[185]
[186]

0.94

1
2

[185]
[145]
[147]
[186]

Pb2V2O7

gr

560
575

[186]
Sr2V2O7

dr
(ppm)

1
2

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

153

Table 5 (continued)
Compound

Ref.

a-NaVO3

[34]
[23]
[185]
[144]
[145]
[189]

KVO3

RbVO3

CsVO3

TlVO3

Cs2V6O16

0
8

31
25

100

28

891

112

21

16

793
794
796

114
106

21
30

4
9

177

48




65
61

90
56

52

0.65
0.87
0.65
0.68
0.67
0.68

582(573)

368

530

820

577
572.7
572.8
569.7

360
355
365
395

530
531
532
536

840
832
822
778

0.63
0.88
0.64
0.65
0.66

547

294

490

856

0.80
0.85

308
227
313
290
307

553
552.7
547.6

300
313
293

500
502
495

870
843
855

0.65
0.14
0.69

313

508

863

0.72
0.64

302
216
314

570(561)

4.33
4.23

565.4

300

517

879

291

0.66

583(571)

330

522

863

3.92
4.1

0.62
0.48

314

0.67

577.4

315

526

3.67
3.76

0.71
0.71

265
265
267

0.72
0.76
0.76

528
529.1
529.1

300
296
294

490
497
497

4.2
2.4

0.55
0.92

512
477

0.17
0.02

510.4
508.4

211
265

298
275

1022
985

255
270

1
2

2.75
2.325

0.98
0.34

470
401

0.01
0.05

509.7
546.0

272
335

277
356

980
947

275
346

1
2

2.45
3.03

0.44
0.89

432
405
459

0
0
0

503
548.1
510.1

290
346
281

290
346
281

935
953
969

290
346
281

1
2

2.664
2.34

0.986
0.30

432
448
400

0.00
0.16
0.05

503
513.8
547.0

290
254
337

290
326
357

935
962
947

290
290
347

1
2

1.25
1.68

0.42
0.69

445
460
400

0.00
0.03
0.05

508
513.0
575.0

296
276
365

296
290
385

953
973
975

296
283
375

0.46
0.49
0.95

4.36

0.75

4.2
4.15

[34]
[26]
[145]
[34]
[24]
[145]

[34]
[23]
[185]
[144]
[145]

[34]
[144]
[145]

[34]
[149]

90
102

247
253
263
259
249
208

3.8
3.77
2.3

[34]
[149]

15
75

d33
(ppm)

VO5,6 axial symmetry


b-NaVO3
[144]
[189]

Rb2V6O16

15
3

d22
(ppm)

0.6

[34]
[144]

30
156

d11
(ppm)

3.65

K2V6O16

c/o

diso
(ppm)

gQ

[195]

b/o

gr

CQ
(MHz)

(NH4)2V6O16

a/o

dr
(ppm)

Site

d^
(ppm)

Tl2V6O16

[34]

465

0.00

700

485

485

1165

485

a-Mg(VO3)2

[34]
[26]
[146]

373
220
310

0.43
0.00
0.30

577

310

470

950

533.9

332

425

844

509

0.15

571

278

355

1080

317

Ca(VO3)2

6.79
7.50

[34]
[24]
[19]
[26]
[146]
[189]

0.63
0.34

3.3
3.16
2.81
3.06
2.94

0.8
0.6
0.6
0.51
0.67

517
530

0.18
0.13

563.0
556.3

258
257

351
326

1080
1086

305
292

74

86

35

Ca(VO3)24H2O

[146]

1
2

4.18
3.73

0.92
0.57

297
492

0.30
0.16

580.0
529.5

387
244

476
323

877
1022

284

150
144

34
50

54
88

a-Sr(VO3)2

[146]

1
2

4.22
5.65

0.12
0.31

218
244

0.32
0.61

639.1
585.9

495
389

565
538

857
830

128
56

86
6

43
54

Ba(VO3)2

[34]
[146]

1
2

3.68
5.56

0.16
0.04

249
190
265

0.30
0.41
0.59

660(701)
658.5
590.6

540
525
380

614
602
536

950
849
856

153
20

70
8

60
90

6.86
4.84

0.40
0.27

387
333
256

0.36
0.02
0.49

517(533)
493.8
491.4

270
324
301

410
331
426

920
827
747

328

53

330

0.18

500

305

365

830

335

484

0.15

521.5

243

316

1006

280

90

47

Zn(VO3)2

[185]
[146]
[189]

Cd(VO3)2

[34]

a-Cd(VO3)2

[146]

1.70

1.00

b-Cd(VO3)2

[146]

6.46

0.47

311

0.31

468.2

264

361

779

22

55

Pb(VO3)2

[185]
[146]

3.75
6.98

0.13
0.31

457
465
428

0.02
0.15
0.12

533(543)
529.8
479.7

310
262
240

320
332
291

1000
995
908

315
297
266

90
90

36
50

0
0

0.797

0.00

640
645

0
0.11

610
609.0

310
251

310
322

1270
1254

310
287

42

126

V2O5

[34]
[93]

1
2

(continued on next page)

154

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

Table 5 (continued)
Compound

Ref.

Site

[148]

d11
(ppm)

d22
(ppm)

d33
(ppm)

d^
(ppm)

0.20
0.15
0.11

609
612

252
259

340
329

1229
1248

296
294

574
611
613
640
787
667
591
1056
730
713

0.49
0.51
0.00
0.00
0.62
0.30
0.15
0.81
0.53
0.14

622.5
618.5
619.4
619.4
593.2
592.6
594.6
597.2
663.6
572.3

194
157
313
299
45
159
255
358
105
166

476
469
313
299
444
359
343
497
492
266

1197
1230
1233
1259
1380
1259
1186
1653
1394
1285

575
780
460
600

0.50
0.60
0.53
0.14

623
596
662
576

377
249
463
358

541
509
601
412

952
1029
923
958

758

0.049

617(614)

218

255

1370

gQ

dr
(ppm)

gr

0.805
0.811
0.799

0.04
0.04
0.21

593
620
636

1
10
100
1000
2
20
200
2000
3
4

1.34
1.33
0.83
0.86
1.39
1.71
1.70
1.61
0.76
0.73

0.7
0.45
0.08
0.10
0.60
1.00
1.00
0.98
0.94
0.31

1
2,4
3
5

1.34
1.40
1.40
0.73

0.70
0.60
0.94
0.30

[19]
[196]
[46]
V2O5nH2O
xerogel

diso
(ppm)

CQ
(MHz)

a/o

b/o

c/o

90
58
140

142
128
127

180

145

313
299

299

216

V2O51.5H2O
xerogel

[197]

VOAsO4

[34]

VOPO4

[34]

841

0.00

734(705)

285

285

1547

285

aI-VOPO4

[180]

1.55

0.55

820

0.00

691.0

281

281

1511

281

54

163

b-VOPO4

[180]
[198]

1.99
1.45

0.59
0.44

818
818

0.00
0.05

755.0
735

346
306

346
346

1573
1553

346
326

43
35

15

aII-VOPO4

[180]
[198]

0.825
0.625

0.52
0.09

582
922

0.67
0.08

776.0
755

287
256

680
330

1358
1680

293

7
0

0
3

60
0

0.554
1.32

0.68
0.55

955
942

0.15
0.07

755
739

206
235

349
301

1710
1681

278
268

28
81

42
77

330

1
2

237

c-VOPO4

[198]

K2V8O21

[34]

480

0.00

570

330

330

1050

KVO3H2O

[34]

604

0.00

606

305

305

1210

305

NaVO3H2O

[34]
[25]

500

0.00

530

280

280

1030

280

NaVO31.5H2O

[197]

420
370

0.00
0.00

548
535

408
412

408
412

828
782

408
412

VO6 isolated
NH4VO(SO4)2

[34]

902

0.04

658

187

227

1560

207

NaVO(SO4)2

[34]

838

0.03

592

160

185

1430

173

KVO(SO4)2

[34]

900

0.04

650

180

220

1550

200

RbVO(SO4)2

[34]

888

0.04

662

198

235

1550

217

CsVO(SO4)2

[34]

859

0.05

671

220

263

1530

242

KVO(SeO4)2

[34]

850

0.05

630

185

226

1480

206

RbVO(SeO4)2

[34]

834

0.04

636

203

235

1470

219

CsVO(SeO4)2

[34]

795

0.04

575

160

195

1370

178

KVOSO4SeO4

[34]

870

0.03

630

180

210

1500

195

RbVOSO4SeO4

[34]

864

0.03

636

190

220

1500

205

VO6 dimer
K4(VO2)2(SO4)2S2O7

[34]

704

0.06

696

325

365

1400

345

Rb4(VO2)2(SO4)2S2O7

[34]

722

0.08

703

315

371

1425

343

Cs4(VO2)2(SO4)2S2O7

[34]
[199]

650
641.5
641.5

0.00
0.07
0.07

690
671.5
678.5

365
326
333

365
374
381

1340
1313
1320

365
350
357

1
2

1
2

3.94

0.64

3.10
2.40

0.90
0.80

2.97
2.97

0.74
0.74

VO6 associated
K3VO2(SO4)2

[34]

494

0.22

566

266

374

1060

Rb3VO2(SO4)2

[34]

626

0.15

562

201

297

1188

249

Cs3VO2(SO4)2

[34]

621

0.15

559

201

297

1180

249

Na3VO2SO4S2O7

[34]

737

0.03

578

200

220

1315

210

K3VO2SO4S2O7

[34]

660

0.13

600

230

315

1260

273

V2O3(SO4)2

[34]

778

0.11

632

204

287

1410

246

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

155

Table 5 (continued)
dr
(ppm)

gr

diso
(ppm)

d11
(ppm)

d22
(ppm)

d33
(ppm)

d^
(ppm)

[34]

697

0.11

633

247

322

1330

285

[34]

664

0.05

696

350

380

1360

365

RbVO2SO4

[34]

709

0.09

711

325

390

1420

358

CsVO2SO4

[34]

466

0.12

544

284

340

1010

312

KVO2SeO4

[34]

520

0.00

540

280

280

1060

280

RbVO2SeO4

[34]

428

0.00

572

358

358

1000

358

CsVO2SeO4

[34]

465

0.19

535

258

348

1000

303

KVO2SO43H2O

[34]

517

0.21

503

190

300

1020

RbVO2SO43H2O

[34]

513

0.23

507

190

310

1020

Compound

Ref.

V2O3(SeO4)2
KVO2SO4

Site

CQ
(MHz)

gQ

KVO2SeO43H2O

[34]

488

0.27

507

197

330

995

RbVO2SeO43H2O

[34]

503

0.28

492

170

310

995

CsVO2SeO43H2O

[34]

498

0.27

500

183

318

998

NaK2[(VO2)3(SO4)3]5H2O

[34]

460

0.00

490

260

260

950

260

NaRb2[(VO2)3(SO4)3]5H2O

[34]

446

0.00

494

270

270

940

270

NaCs2[(VO2)3(SO4)3]5H2O

[34]

447

0.00

503

280

280

950

280

a/o

b/o

c/o

VO6 isolated
[(n-C4H9)4N]3[VW5O19]

[200]

0. 605

0.65

200

0.95

504.8

310

500

705

90

30

Cs3[VW5O19]

[200]

1.30

0.8

466

0.10

519.6

263

310

986

287

70

40

K4PVW11O40

[201]

0.94

0.48

537

0.16

561.3

250

336

1097

293

64

26

[(n-C4H9)4N]4PVW11O40

[201]

1.13

0.61

521

0.13

543.0

249

317

1064

283

52

42

Cs4PVW11O40

[201]

1.25

0.53

589

0.25

563.9

196

343

1153

269

55

52

[NaxCs4x]PVW11O40

[201]

1.15

0.15

590

562.4

267

267

1152

267

63

50

a1-4-K7P2VW17O62

[124]

1.73

0.37

a2-4-K7P2VW17O62

[124]

0.88

0.75

50

VO6 dimer
[(n-C4H9)4N]3H[V2W4O19]

[200]

1.05

0.95

418

0.10

505.2

275

317

923

296

70

Na2Cs2[V2W4O19]6H2O

[200]

1.56

0.35

456

0.20

526.0

252

344

982

298

80

60

50

a-1,2-H3K2[PV2W10O40]
CH3OH14H2O

[201]

1.79

0.81

516

0.07

542. 5

266

303

1059

285

62

36

a-1,2-[(n-C4H9)4N]5PV2W10O4

[201]

1.75

0.77

549

0.01

557.0

280

285

1106

283

81

39

a-1,2-Cs5PV2W10O40

[201]

1.51

0.57

479

0.0

533.1

294

294

1012

294

63

42

VO6 three and more associated


a-1,2,3-K6PV3W9O40

[201]

3.93

0.87

371

0.0

522.1

337

337

893

337

38

a-1,2,3-[(n-C4H9)4N]6
PV3W9O40

[201]

5.50

0.75

451

0.05

444.4

208

230

895

219

45

40

a-1,2,3-Cs6PV3W9O40

[201]

3.5

1.0

428

0.0

519.9

306

306

948

306

87

Cs4[H2V10O28]4H2O

[202]

1
2
3
4
5

3.0
2.5
2.5
4.0
6.4

0.9
0.7
0.9
0.5
0.4

400
600
450
450
350

0.20
0.70
0.50
0.20
0.20

518.0
531.0
540.0
545.0
427.0

278
21
203
275
217

358
441
428
365
287

918
1131
990
995
777

318

320
252

Yellow foam

[202]

1
2
3

3.90
3.00
2.00

0.30
0.90
1.00

400
350
140

0.20
0.20
0.20

435.0
530.0
564.0

195
320
480

275
390
508

835
880
704

235
355
494

White foam

[202]

1
2

2.00
2.50

1.00
0.70

100
200

0.20
0.30

564.0
578.0

504
448

524
508

664
778

514
478

Green foam

[202]

3.00

0.90

350

0.50

525.0

263

438

875

K5NaV10O2810H2O

This work

1
2

3.70
2.20

0.90
0.99

300
345

0.10
0.07

424.0
488.0

259
303

289
328

724
833

274
316

K4Na2V10O2810H2O

This work

1
2
3
4

3.70

0.90

300

0.10

262

292

727

277

2.20

0.99

345

0.07

304

329

834

317

421.5
426.8
431.0
489.0

(continued on next page)

156

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

Table 5 (continued)
Compound

Ref.

Site

CQ
(MHz)

gQ

dr
(ppm)

diso
(ppm)

gr

5
6
7
8
Rb4Na2V10O2810H2O

(NH4)4Na2V10O2810H2O

Cs2V4O11

This work

This work

[149]
[34]

Bi4V2O11

[189]

1
2
3
4
5
6
7
8

[149]
[34]

d22
(ppm)

d33
(ppm)

d^
(ppm)

259

289

724

274

304

329

834

317

258

288

723

273

305

330

835

318

3.70

0.90

300

0.10

2.20

0.99

345

0.07

3.70

0.90

300

0.10

2.20

0.99

345

0.07

1
2
1
2

1.25
1.68

0.42
0.69

620
256
610
253

0.06
0.13
0.05
0

510
575;581
510
575(572)

184
430
190
445

222
464
220
445

1133
831
1120
825

203
447
205
445

1
2

4.0
4.5

0.98
0.96

48
489

0.20
0.08

498.6
422.2
510

470
158

480
197

546
911

475
178

1
2
1
2

1.75
0.22

0.40
0.20

480
80
486

0.08
0.60
0.00

517.0
624.0
500(503)
620

258
560
260

296
608
260

997
704
990

277

260

1
2
3
4
5
6
7
8

424.0
430.0
435.0
489.0
493.0
497.0
505.0
511.0
423.0
427.0
432.0
490.0
495.0
502.0
508.0
514.0

Rb3V5O14

[34]

1
2

463

0.00

496(517)
619

285

285

980

285

Tl3V5O14

[34]

1
2

489

0.00

500(504)
594
508

260

260

993

260

Pb5(VO4)3Cl

[185]

VO[SiO(OtBu)3]3

[149]

VO[SiOPh3]3

[177]

[(c-C6H11)7(Si7O12)VO]2

[177]

VO[OAmt]3

[204]

VO[OSiMe3]3

[204]

(SiO)3VO

[177]

(SiO)3VO2H2O

[177]

a/o

b/o

c/o

90

495.0
501.0
506.0
511.0

[203]
K3V5O14

d11
(ppm)

2.45

0.1

445

0.05

776

542

564

1221

553

3.19
3a

0.13

414
420
401

0.04
0.08
0.087

736
731.4
714

510
504
500

525
538
535

1150
1151
1115

518
521
518

107

0.15

681

619

636

788

628

2.5a

204

0.05

714

607

617

918

612

480

0.042

710

450

470

1190

460

620

0.129

580

250

330

1200

290

Values of CQ estimated in this work.

Negative values of dr, detected in some thortveitite compounds, as well as for one of the vanadium sites in aMg2V2O7, is explained by certain structural features of
the local vanadium coordination. For example, in bMg2V2O7 besides four VO bonds in VO4 tetrahedra there
, which can be
is an additional VO distance at 2.42.8 A
described as the fth VO bond.
According to Hawthorne and Calvo [165] in barium pirovanadate Ba2V2O7 there are four structurally non-equivalent
types of VO4 tetrahedra. Similar ndings were also reported
for isostructural Sr2V2O7 [166]. All four of these VO4 tetrahedra are distorted in a somewhat similar fashion, while three
VOt distances remain almost equal in length, a bridging
VOl2 bond forming a V2 O7 4 dimer is usually longer by

. This type of distortion results in 51V MAS


about 0.1 A
NMR spectra illustrated in Fig. 21 for Ba2V2O7. The MAS
spinning sidebands from the central and satellite transitions
extend over 10,000 ppm (1 MHz at 9.4 T). At the same time
the central transition is suciently well resolved to show
three individual lines with relative intensities of 1:2:1. Since
according to the structural data there are four non-equivalent
vanadium sites in the crystal structure, it is clear that two of
these sites overlap in this spectrum. The 51V NMR parameters determined for the four sites are very close (Table 5).
Summarizing 51V NMR data available for Q1 vanadium
sites (Table 5), we can conclude that vanadium in slightly
distorted tetrahedral sites sharing one common oxygen
atom with the adjacent tetrahedra (Q1 type) has larger

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

157

-560

-580

-600

-620

, ppm

3000

2000

1000

-1000

-2000

-3000

-4000

-5000

, ppm
Fig. 21. 51V MAS NMR spectrum of Ba2V2O7 obtained at 9.4 T using 10 kHz MAS. (1) Full spectrum. (2) Isotropic region of the MAS spectrum showing
three isotropic lines.

CSA anisotropy, 100 < dr < 200 ppm, compared with Q0


sites; the quadrupolar coupling constant varies from 2.5
to 10 MHz, while the CSA asymmetry parameter gr
changes from 0.1 to 0.9 [93,94,149].
5.1.3. Tetrahedral Q2 sites
Tetrahedral sites of Q2 type are found in metavanadates
of ammonia, alkali metals and some other monovalent
metals, MVO3, (M+ = Li [167], Na [168], K and NH4
[169], Tl [170], Rb, and Cs [171]). In these compounds corner-sharing VO4 tetrahedra form innite two-dimensional
chains. As a result, the VOl2V bonds are further elongated, thus causing additional distortions of VO4 tetrahedra. In many MVO3 compounds with Q2 sites 51V NMR
parameters have been carefully characterized with
SATRAS [144] and were found to fall into quite narrow
ranges. For example, the quadrupolar coupling constant
in MVO3 is often within 2.84.3 MHz, while the chemical
shielding anisotropy is within the 217314 ppm range.
The crystal structure of Ba(VO3)2 is similar to metavanadates of alkali metals [172], where corner-sharing VO4 tetrahedra form extended 2D-chains of Q2 units, although in
Ba(VO3)2 there are two non-equivalent vanadium sites.
Generally speaking, vanadium in strongly distorted Q2
tetrahedral sites with adjacent tetrahedra sharing two common oxygen atoms has large values of anisotropy
(200 < dr < 500 ppm); the quadrupolar coupling constant
varies from 2 to 7 MHz; and the CSA asymmetry gr ranges
from 0.6 to 0.8 [93,94,149].
5.1.4. Associated non-axial VO5 and VO6 sites
Associated non-axial VO5 and VO6 sites are typical for
some metavanadates of divalent metals M(VO3)2. For

example, distorted octahedral VO6 sites are found in compounds of the brannerite structural type (analogs of
ThTi2O6) with space group C2/m, i.e. in Mg(VO3)2 [173],
Zn(VO3)2 [174] and b-Cd(VO3)2 [175]. The low-temperature modication a-Sr(VO3)2 (Pnma) contains two nonequivalent distorted VO6 sites [176]. 51V NMR parameters
for these compounds were reported in [146] and are summarized in Table 5. 51V NMR spectra are characterized
by CSA values similar to those for Q2 sites,
(200 < dr < 400 ppm), however, the CSA asymmetry
parameter gr is smaller that in Q2, 0.3 < gr < 0.6, but somewhat higher than in pyramidal vanadium sites (see below).
5.1.5. Isolated and associated trigonal VO4 pyramids
Trigonal VO4 pyramids are usual in VO[SiO(OtBu)3]3,
VO[SiOPh3]3, [(c-C6H11)7(Si7O12)VO]2, (SiO)3VO, and
similar systems. Vanadium in these compounds is fourcoordinated with one short V@O bond. When VOl2V
bonding is absent or very weak, the pyramids are said to
be isolated. Vanadium occupying these sites has an axially
symmetric chemical shielding tensor with a large anisotropy (400 < dr < 500 ppm), the quadrupolar coupling constant can vary from 1 to 4 MHz, while the CSA asymmetry
parameter gr changes from 0 to 0.2, and d^  400 to
600 ppm [144149,177].
Pyramidal association, i.e. as observed in Cs4V2O11,
leads to smaller anisotropy mostly due to changes in dII,
while d^ remains practically unchanged (at 450 ppm in
Cs4V2O11). The crystal structure of Cs2V4O11 has been
reported in [178]. There are three dierent vanadium sites
in this structure, including two types of trigonal VO4 pyramids, and one type of a trigonal VO5 bipyramid having different degrees of association (Fig. 22). Corresponding 51V

158

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

V2 V3

V1

-500

, ppm

-600

3
1500

500

-500
, ppm

-1500

-2500

Fig. 22. 51V MAS NMR spectra of Cs2V4O11 obtained at 9.4 T. (1) Experimental 51V MAS spectrum. (2) Simulated spectrum corresponding to VO4 sites.
(3) Dierence spectrum corresponding to VO5 sites. Inset shows the isotropic region of the MAS spectrum. The crystal structure of Cs2V4O11 is given on
the left.

5.1.6. Isolated octahedral VO6 and tetragonal VO5 pyramids


51
V NMR parameters for four of ve known [179] crystalline modications of VOPO4 are summarized in Table 5.
In VOPO4 vanadium occupies distorted octahedral VO6
sites. Each of these octahedrons is composed of four oxygen atoms in equatorial positions and two axial oxygen
atoms. One of two axial oxygens has a short V@O bond,
therefore the vanadium coordination in such sites is

MAS NMR spectra for this compound are shown in


Fig. 22. Three isotropic lines can be easily identied in
these spectra, two narrower lines with close isotropic chemical shifts belonging to trigonal VO4 pyramids, and a
broader line attributable to vanadium in distorted octahedral environment of trigonal VO5 bipyramids. 51V NMR
parameters determined for all three sites are given in
Table 5.

*
2

1000

-1000
, ppm

-2000

-3000

Fig. 23. 51V NMR spectra of aI-VOPO4 at 9.4 T. (1) Experimental static spectrum. (2) Experimental 10 kHz MAS spectrum. (3) Simulated MAS spectrum
calculated with the following parameters, CQ = 1.55 MHz, gQ = 0.55, dr = 880 ppm, gr = 0, diso = 691 ppm. The building unit of the aI-VOPO4 is shown
above. The isotropic line is marked with asterisk.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

159

B
C

-500

-1000

-1500

Fig. 24. 51V NMR spectra of Cs4(VO)2O(SO4)4 measured at 9.4 T. (A) 14 kHz MAS spectrum. (B) Isotropic region of the MAS spectrum. The isotropic
lines are marked with asterisks. (C) Static spectrum.

Vanadium in isolated octahedral pyramids has an axially


symmetric chemical shielding tensor with a large anisotropy
(750 < dr < 900 ppm); at the same time the quadrupolar coupling constant is often less than 2 MHz, while gr changes
from 0 to 0.2, and d^  180 to 250 ppm.
Association of octahedral pyramids is usually accompanied by some decrease in anisotropy, as was observed, for
example, for dimeric species in Cs4(VO)2O(SO4)4 where
dr is about 700 ppm (Fig. 24).

described as octahedral pyramidal. Since there is no bridging via VOl2V corner sharing, individual VO6 octahedra
are considered isolated. Coordination of vanadium sites in
aI-VOPO4 together with the corresponding 51V NMR spectra is shown in Fig. 23 [180].
Similar isolated octahedral VO6 pyramids are also typical for VOAsO4, and some other compounds including
MVO(SO4)2, (M+ = NH4, Na, K, Cs, Rb), MVO(SeO4)2
(M+ = K, Rb, Cs), and MVO(SO4)(SeO4) (M+ = K, Rb).
A

B
1

-400

-500 -600
ppm

2
1

2000

-2000
, ppm

-4000

2000

-2000

-4000

, ppm

Fig. 25. 51V MAS NMR spectra of Rb2V6O16 recorded at 9.4 T using 14 kHz MAS. (A1) Experimental spectrum. (A2) Simulated spectrum composed of
two sub-spectra. (B1) First sub-spectrum calculated with CQ = 2.66 MHz, gQ = 0.99, dr = 448 ppm, gr = 0.16, diso = 514 ppm. (B2) Second subspectrum calculated with CQ = 2.34 MHz, gQ = 0.30, dr = 400 ppm, gr = 0.05, diso = 547 ppm. (C) Isotropic region of the MAS spectrum with asterisks
indicating isotropic lines for two vanadium sites.

160

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

5.1.7. Associated tetragonal pyramids


Association of tetragonal VO5 pyramids occurs in
M2V6O16 (M = NH4, Cs, Rb, Tl). All these compounds
are isostructural with the P21/m space group [181]. There
are two types of vanadium site in this structure occupying
two types of tetragonal pyramids in the ratio of 1:2. Each
pyramid is connected to two adjacent pyramids via edge
or corner sharing.
The 51V MAS NMR spectrum of Rb2V6O16 shown in
Fig. 25 indicates the presence of two non-equivalent vanadium sites in the ratio of 1:1.8, in good agreement with the
crystal structure described above. The 51V NMR parameters for these two sites do not dier signicantly (Table
5). Some metavanadates of divalent metals also have associated tetragonal VO5 pyramids in their structure, these
include, for example, Ca(VO3)2 and a-Zr(VO3)2 (Table 5).
5.1.8. Strongly associated octahedral sites in decavanadates
The molecular structure of a V10 O28 4 polyanion was
rst reported by Evance [182]. The model structure of this
polyanion is presented in Fig. 26A. (adopted from Ref.
[183]). This structure consists of 10 VO6 octahedra labeled
as V1 ; V1  ; V2 ; V2  ; V3 ; V3  ; V4 ; V4  ; V5 ; and V5  , which
within the polyanion results in only three non-equivalent
vanadium sites in the ratio of V5/V3,4/V1,2 = 2:4:4. Two
V5 octahedra are buried entirely inside the polyanion,
and their edges are shared with other octahedra in the
structure. Two pairs of V3,4 and V1,2 octahedra are similar

in their structure and form an external part of the polyanion. In aqueous solutions, V10 O28 6 anions are stable and
show characteristic 51V NMR chemical shifts of
425 ppm for V5, 506 ppm for V3,4, and 524 ppm for
V1,2 sites [184].
In the solid state all vanadium sites in the polyanion
become non-equivalent, as observed, for example, in
K5NaV10O2810H2O, K4Na2V10O2810H2O, Rb4Na2V10O28
10H2O, and (NH4)4Na2V10O2810H2O. As a result, corresponding 51V MAS NMR spectra are usually very complicated due to closely overlapping lines from dierent sites.
The isotropic shifts are usually in the range of 425 to
500 ppm. While the magnetic shielding anisotropy is
rather small for all these sites, it is smallest for internal
V5 octahedra with the highest degree of association.
The nature of charge-balancing cations in decavanadates has only moderate eects on the isotropic shift values. The resonances corresponding to V5 sites experience
shift from 425 to 430 ppm upon changing cations from
K+ to NH4 and nally to Rb+. The V3,4 and V1,2 resonances at around 500 ppm become more separated for
NH4 , and somewhat less separated for Rb+.
5.2. Correlating local environment of vanadium nuclei in
VOx species with 51V NMR parameters
Studies of vanadium-51 NMR spectroscopy of solid
samples has a long tradition. There is a large volume of

*
*

2000

-2000

-4000

2000

, ppm

3000

2000

1000

-1000

-2000

-3000

-2000

, ppm

-4000

-4000

-5000

, ppm
Fig. 26. 51V 10 kHz MAS NMR spectra of decavanadates obtained at 9.4 T. (A) Structure of V10 O28 6 from Ref. [183]. (B) K5NaV10O2810H2O. (C)
K4Na2V10O2810H2O. (D) Rb4Na2V10O2810H2O. The isotropic lines are indicated with asterisks. In (C and D) the experimental spectra are compared
with the simulated spectra shown below. The calculation parameters used in simulations were as following: (C2) CQ = 3.7 MHz, gQ = 0.9, dr = 300 ppm,
gr = 0.1, diso = 424 ppm; (D2) CQ = 2.2 MHz, gQ = 0.99, dr = 345 ppm, gr = 0.07, diso = 489 ppm.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

161

for >200 ppm, <0.2

1.0
4

0.8
VO4 trigonal pyramids

CQ, MHz

0.6

0.4

VO6 strongly associated


VO5,6 associated pyramids
VO5,6 isolated pyramids

0.2
0
-200

0
0

200

400

600
, ppm

800

1000

VO5 axial
VO6 axial

-400
, ppm

-500

-600

Fig. 28. Correlation between the value of perpendicular component of the


51
V NMR chemical shielding tensor, d^, and the quadrupolar coupling
constant, CQ, for vanadium polyhedra with one short V@O bond
characterized by dr > 200 ppm and gr < 0.2.

VO4 axial

-300

VO4 Q

VO4 Q1
VO4 Q2
VO5,6 non-axial

51

Fig. 27. Correlation between V asymmetry, gr, and anisotropy, dr,


parameters of chemical shielding tensor obtained for various vanadium
compounds.

experimental data available for a wide variety of systems


and materials. In Table 5 we summarize all the experimental results currently available in the literature and some
results just recently obtained by us for individual vanadium
oxide compounds with well-characterized vanadium local
environment. These results have been accumulated over
years by many research groups often working with dierent
conventions for reporting CSA and quadrupolar parameters. For convenience of use, all the data listed in Table 5
have been recalculated from those originally reported and
are now presented according to notations adopted in this
review as outlined above.
There have been several reported attempts to correlate
the local structure of vanadium sites with 51V NMR
parameters. The most comprehensive correlation reported
earlier by Lapina et al. [34] was between the coordination
of vanadium in VOx polyhedra and the value of the chemical shift anisotropy. Using this correlation these authors

Table 6
Eight types of VOx sites that can be recognized by
0

51

were able to distinguish four dierent types of vanadium


environments, including tetrahedral sites Q0, Q1, and Q2,
as well as vanadium in distorted octahedral environment
of the V2O5 type.
In this report, we are revising these earlier observations
by specifying a much wider range of vanadium environments. Correlating the chemical shielding anisotropy and
the CSA asymmetry parameter as shown in Fig. 27 it is
now possible to clearly dierentiate ve dierent types of
vanadium coordination environments. In addition to tetrahedral sites Q0, Q1, Q2 these now include non-axial VO5
and VO6 sites, and axial VOn (n = 4, 5, 6) species. The latter
group contains isolated and associated VO6 octahedra, isolated and associated VO5 tetragonal pyramids, and trigonal
VO4 pyramids, which are dicult to discriminate using
only CSA parameters.
In addition to characteristic values of the chemical
shielding anisotropy and the CSA asymmetry parameters
for each site outlined in Table 6, we also note several other
useful tendencies and correlations.
The rst observation is related to the value of the isotropic chemical shift, which in Q0 sites was found to depend
on the nature of the counter-ion. Thus, the isotropic chemical shift will usually increase as following: M+ = Li

V NMR

Tetrahedral Q sites
Tetrahedral Q1 sites
Tetrahedral Q2 sites
Associated non-axial VO5 and VO6 sites
Isolated octahedral VO6 and tetragonal VO5 pyramids
Associated pyramids VO5 and VO6
Strongly associated octahedra VO6
Trigonal pyramids VO4

dr (ppm)

gr

<100
100200
200500
200400
>700
200400
300400
400500

0.10.9
0.60.8
0.30.6
00.1
00.2
00.2
00.2

d^ (ppm)

CQ (MHz)

Key NMR parameters

200350
200400
200350
400500

06
2.510
27
3.56
02
23
34
14

dr,
dr,
dr,
dr,
dr,
dr,
dr,
dr,

gr
gr
gr
gr
gr,
gr,
gr,
gr,

d^,
d^,
d^,
d^,

CQ
CQ
CQ
CQ

162

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

Na K Cs; M2+ = Zn Mg Ba Sr; M3+ = Bi


Ce La Lu Y Sc.
In a similar fashion, the quadrupolar coupling constant,
CQ, in vanadates of M+ is typically ranging from 1 to
3.5 MHz, in M2+ vanadates this constant falls in a very
narrow range of 0.51 MHz, and in M3+ vanadates CQ
can be from 3.5 to 5.6 MHz. Such a broad distribution in
CQ is most likely due to relatively short VO distances in
. Even small deviavanadates normally not exceeding 1.7 A
tions of these distances from the average value would result
in considerable electric eld gradients at the location of
vanadium nucleus.
The chemical shift anisotropy in Q0 sites does not exceed
100 ppm, indicating high symmetry of the vanadium coordination in these species.
Unfortunately, the values of the chemical shift anisotropy and the CSA asymmetry parameter (Fig. 27) do not
allow dierentiating between various types of axial vanadium species, including isolated and associated VO6 octahedra, isolated and associated VO5 tetragonal pyramids,
and trigonal VO4 pyramids. This can be done, however,
by analyzing eective values of d^ and the value of the
quadrupolar coupling constant in cases with large CSA,
dr > 200 ppm, and gr < 0.2. The eective values of d^ can
be estimated as d^  1/2 (d1 + d2), where di are the components of the CSA tensor. An example of such correlation is
shown in Fig. 28. Using this correlation it is now possible
to identify four more types of vanadium coordination with
axial symmetry of CSA (Table 6).
We also note that while isolated pyramids would usually
demonstrate large magnetic shielding anisotropies, an association of pyramids often leads to decreasing values of the
anisotropy. The situation is opposite for tetrahedra, where
the smallest anisotropy is always found in isolated species,
while it is gradually increasing with the extent of association as shown in Fig. 29.

11

22

chains

33

||

3D structure

dimer

2D structure

isolated

isolated

-500

-1000

ppm

-500 -1000

Fig. 29. Correlation between local structure of vanadium sites and


association of VOx polyhedra with the type and magnitude of CSA. The
value of di is determined by a V@O distance: variation of this distance
leads to di variation from 1600 to 800 ppm. The value
from 1.56 to 1.68 A
of d^ corresponds to dierent types of VOx pyramids: for VO5 and VO6 it
lies in the range from 200 to 350 ppm, for VO4 in the range 450 to
600 ppm.

Using results presented above it is now possible to dierentiate up to eight dierent vanadium coordination environments by their 51V NMR parameters. These
correlations will be useful for studying new vanadium
oxide systems, in particularly those systems and materials
with disordered or poor crystalline structures.
6.

93

Nb NMR data compilation

6.1. Chemical shielding and quadrupolar tensor parameters


in individual niobium compounds
The niobium coordination number in niobium-containing oxide compounds varies from 4 to 8.
6.1.1. Six-coordinated compounds
The most typical are compounds with six-coordinated
niobium. In this work the 93Nb NMR parameters are summarized for the following six-coordinated niobium compounds: Li3NbO4 [205,206], Bi3NbO7 [207], La3NbO7
[208], BiNbO4 [209], LiNbO3 [210], NaNbO3, KNbO3
[211,212], SnNb2O6 [213], Sn2Nb2O7, K8Nb6O19, Te3Nb2O11
[214], NbVO5 [163], VNb9O25 [164], Pb(Mg1/3Nb2/3)3
(cubic, tetragonal, and rhombic symmetry), Pb3Nb4O13,
PbNb2O6, Pb2Nb2O7, Pb5Nb4O15, Pb3Nb2O8 [47]. 93Nb
NMR parameters for these compounds are summarized
in Table 7.
Niobium compounds with cubic symmetry, including
Li3NbO4, Pb(Mg1/3Nb2/3)3, Bi3NbO7, and Pb3Nb4O13 have
relatively small quadrupolar coupling constants usually not
exceeding 20 MHz. The 93Nb isotropic chemical shift for
these compounds is found within the 900 to 990 ppm
range. Measurements at high magnetic elds allow determination of the chemical shift anisotropy, which is normally
not exceeding 140 ppm. Because CQ and the chemical shift
anisotropy are not extremely large, it is often possible to
accurately calculate a full set of 93Nb NMR parameters
which ts all the experimental spectra, including static,
MAS, HFMAS, and 3QMAS spectra.
Examples of MAS, HFMAS, and static 93Nb NMR
spectra for Li3NbO4 are shown in Fig. 30. The static
93
Nb NMR spectrum recorded at 97.7 MHz (9.4 T) has
singularities of the rst-order quadrupolar perturbations
(Fig. 30B). Note, that in addition to the central transition
only the 1/2 M 3/2 transitions can be observed under
these experimental conditions. The value of CQ estimated
from the static spectrum is 12 MHz. In the 93Nb MAS
spectra at 9.4 T the satellite transitions for m = 1/2, 3/2,
and 5/2 can also be seen (Fig. 30A). The spinning sidebands analysis described above produced accurate values
of CQ = 11.5 MHz and gQ = 0.1. The 93Nb MAS NMR
spectrum recorded at higher eld had revealed a moderate
chemical shift anisotropy, dr = 140 ppm (Fig. 30C). The
isotropic chemical shift for Li3NbO4 was found at
950 ppm in all these experiments.
93
Nb NMR spectra of six-coordinated compounds with
non-cubic symmetry, i.e. orthorhombic, rhombohedral,

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

163

Table 7
93
Nb NMR parameters of individual niobium compounds
Compound
Four-coordinated
YNbO4

LaNbO4
PrNbO4
Five-coordinated
Na5NbO5
CaNb2O6
Six-coordinated
Pb(Mg1/3Nb2/3)O3
Bi3NbO7
Li3NbO4

a-BiNbO4

Pb3Nb4O13
Pb(Mg1/3Nb2/3)O3
Pb(Mg1/3Nb2/3)O3
La3NbO7
K8Nb6O19
SnNb2O6
LiNbO3

Cs4Nb11O30 (3 sites)
NaNbO3
NaBa2Nb5O15 (2 sites)
KNbO3
Sn2Nb2O7
PMN pyrochlore
PbNb2O6
Te3Nb2O11

PNb9O25 (3 sites)

Ref.

CQ (MHz)

gQ

[55]
[55]
This work
[55]
This work
[55]
This work

81
80
82
70
89
87
87

0.41
0.5
0.38
0.3
0.15
0.25
0.25

[55]
[55]
[55]
[55]

11.1
11.1
50
50

0.01
0.01
0.8
0.8

[47]
[55]
[55]
[55]
[55]
[55]
[55]
[55]
This work
[50]
[47]
[47]
[55]
This work
This work
[55]
[55]
[55]
[55]
[35]
[21]
[50]
[222]
[50]
[222]
[222]

<0.8
<20
11.5
12.0
12.0
23
23
20.7
20.7
13.7
17
>62
49
49
86
38.7
40
22
22.2
22
22.0
22.1
16
22.7
20
21
24
21.5
23.1
<20
26.8
16.8
19
22
22

[55]
[47]
[39]
[50]
[50]
[48]
[55]
[55]
[55]
[222]
This work

NbOPO4
NbVO5
Nb3(NbO)2(PO4)7 (3 sites)
Na3.04Nb7P4O29 (4 sites)

[222]
[35]
[222]
[222]

Na4Nb8P4O32 (4 sites)

[222]

Pb2Nb2O7 (9 sites)

[50]

12
25
20
29
68
21
16.5
30
32
27
29
30
13.6
17.0

0.1
0.1
0.1
0.35
0.4
0.79
0.51

0.21
0.275
0.67
0.45
0.01
0.2
0.2
0

0.6
0.80

0.5
0.6
0.6

0.9

diso (ppm)

gr

dr (ppm)

800
750
845
650
800
400
350

903
903
990
990

0
0.6
0.5
0

900
900
950
950
954
963
963
974
977
995
954 to 980
954 to 980
980
968
970
1010
1010
1004
996
1004
1009
1004
1062
1073
1070
940
1079
1015
1069
1050
1014
1113
1090
1166
1176
1207
1177
1202
1167
1204
1250
1316
1338
1583
1287
1328
1285
1338
1003
978

0.48
0
0.3
0

200
0
200
0

0
120
300
100

a, b, c
0, , 0
165, 16, 100

HF
HF

145, 9, 103

HF
HF

0, 31, 90

0.35
0.3
0
0.25

135
140
0
180

0.56

150

24, 22, 78

0
0.69

0
113

50, 27, 72

0
0

Field

HF

HF
HF

HF
HF

0
0

HF
HF
HF
0

230
150
250

Static
MAS
Statica
Static
Statica
Static
Static
MAS
MASa
Static
MAS
Nutation
MAS
MAS
Static
MAS
MAS
MAS
3QMAS
Statica
3QMAS
Nutation
Nutation
Static, MAS
Statica
Static
Static
MAS
Static
MAS
MAS
Single crystal
3QMAS
3QMAS
3QMAS
3QMAS
3QMAS

HF
HF
HF

MAS
MAS
MAS
MAS
3QMAS
Single crystal
MAS
MAS
STMAS
3QMAS

HF

Static and 3QMAS

HF
HF
HF

Static
MAS
MAS
MAS

HF

MAS

HF

0.2
0
0.1

Method

45, 20, 30

3QMAS
(continued on next page)

164

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

Table 7 (continued)
Compound

Ref.

CQ (MHz)

Pb5Nb4O15 (10 sites)

[50]

Pb3Nb2O8 (10 sites)

[50]

16.6
17.9
20.6

H-Nb2O5 (15 sites)

[222]

24

1013
975
999
951
1204

Seven-coordinated
LaNb5O14 (3 sites)

[55]

K2NbF7

[53]

38.5

1200 (NbO6)
1230 (NbO6)
1267 (NbO7)
1600

Eight-coordinated
NaNb3O8 (2 sites)

[55]

Cs3NbO8

This work

diso (ppm)

gQ

0.35

dr (ppm)

gr

a, b, c

Field

Method
3QMAS
3QMAS

200

1250 (NbO7)
1500 (NbO8)
1600

HF

3QMAS

HF

STMAS

HF

MAS
MAS

HF

Static

Simultaneous simulations of MAS and static spectra recorded at 9.4 and 21 T.

A
3/2

1/2

5/2

3/2

3/2

5/2

CT

4000

-920

3800

3500

2500

-960

1500

-1000

500

-2800

-500

-1500

-2500

-3000

-3500

-4500

-5500

, ppm

C
1

2
2000

-2000

-4000

, ppm

-1000

-2000

, ppm

Fig. 30. 93Nb NMR spectra of Li3NbO4 obtained at 9.4 and 21.1 T. Experimental spectra are shown above, calculated spectra are shown below. (A)
10 kHz MAS spectrum at 9.4 T. Details of the central transition +1/2 M 1/2 and two satellite transitions, 3/2 M 1/2 and 5/2 M 3/2 are shown as
insets. (B) Static spectrum obtained with the quadrupolar echo pulse sequence at 9.4 T. (C) 9 kHz MAS spectrum at 21.1 T. The same simulation
parameters were used in all three simulations as following: CQ = 11.5 0.5 MHz, gQ = 0.1, diso = 950 5 ppm, gr = 0.30 0.05, dr = 135 5 ppm.
The two latter parameters were not used in simulations shown in (A).

monoclinic, rhombic, tetragonal, and trigonal, are often


broadened due to strong quadrupolar interactions. Analysis of such spectra is complicated and requires a variety of

approaches as illustrated below for LiNbO3. Of all niobium


compounds, LiNbO3 is the most investigated with solidstate 93Nb NMR [20,21,2729,31,215,216]. As a matter of

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

fact, all new NMR techniques applicable to I = 9/2 have


rst been tested on LiNbO3 [50,65,66,217].
The 9.4 T 93Nb MAS NMR spectrum of LiNbO3 spun
at 30 kHz, shows a lineshape dominated by the secondorder quadrupolar interactions (Fig. 31A). Well-pronounced spinning sidebands from the satellite transitions
1/2 M 3/2, 3/2 M 5/2, and the central transition
+1/2 M 1/2 allows the determination of CQ, gQ, and
diso values (Fig. 31A, Table 7). While improvement is
expected in 93Nb resolution provided by DAS spectroscopy [50], it is oset by the signicant homonuclear
NbNb dipolar interactions dominating the centerband
in the isotropic dimension of the DAS NMR spectrum
of LiNbO3 (Fig. 31B). At the same time, the 93Nb
3QMAS NMR spectrum of LiNbO3 shows resolution
of approximately an order of magnitude better than in
the DAS spectrum (Fig. 31C). The full-widths at a
half-maximum height of the 93Nb resonances of LiNbO3
were found to be 9 kHz in MAS spectra, 5.8 kHz in
DAS, and only 0.7 kHz in a 3Q projection of the
3QMAS spectrum. Since the 93Nb CQ is large in this
case, 22 MHz, the two-dimensional nutation spectroscopy can also provide complementary information
(Fig. 31D). The 2D nutation spectrum of LiNbO3 shows
a single Nb site with its center of gravity at 4mRF corresponding to a CQ value of ca. 20 MHz [48].
Regardless of dierent NMR approaches, either MAS,
Static, DAS, 3QMAS, 2D nutation, or SSTMAS at 9.4 T
(Fig. 31E), the NMR parameters obtained for LiNbO3
are rather similar (Table 7).
Magnetic elds above 1820 T are most benecial for
niobium compounds with large quadrupolar coupling constants exceeding 20 MHz or even larger, as, for example, in
BiNbO4. Static, MAS, 3QMAS at 9.4 T and MAS 93Nb
NMR spectra recorded for this compound at 21 T are
shown in Fig. 32. These spectra are resolved well enough
to accurately measure the magnetic shielding anisotropy,
which in this compound was found to be small (only
150 ppm). Certainly, such small magnetic shielding anisotropy would be very dicult to detect at lower magnetic
elds (Fig. 32A).
At even larger quadrupolar coupling constants, in
excess of 50 MHz, ultrahigh magnetic elds become a
necessity. When static experiments are performed at several elds, at least one high-eld measurement is required
as shown for La3NbO7 (Fig. 33). In this compound the
niobium coordination in zig-zag chains of NbO6 octahedra is very close to symmetric. Nevertheless, due to the
fact that Nb atoms are o-center, the observed 93Nb
NMR line is broad (Fig. 33) and the quadrupolar coupling constant is almost 50 MHz with the symmetry of
the quadrupolar tensor close to axial (gQ = 0.2). The isotropic shift (980 ppm) is somewhat smaller than in alkali
niobates.
The following conclusions can be drawn from analysis
of the six-coordinated niobium compounds with non-cubic
symmetry.

165

(i) In such compounds the isotropic chemical shift seems


to depend on the ionic character of the niobium sublattice. In niobates of M(+1, +2, +3) elements the
niobium sublattice has more anionic character, and
the isotropic shifts are in the range from 1000 to
1100 ppm. When M(+4) or M(+5) is present, the
niobium sublattice has more covalent (towards cationic) character and the isotropic chemical shifts are
shifted to 1200 to 1300 ppm.
(ii) The quadrupolar coupling constant depends on the
site symmetry and not on the coordination number.
CQ can vary in a wide range, from hundreds of kilohertz to hundreds of megahertz.
(iii) The chemical shift anisotropy is more pronounced
when atoms of dierent types are present in the rst
coordination sphere, e.g. O and F.
6.1.2. Four-coordinated compounds
Compounds having Nb with coordination four are
rather rare. A few of them belong to the MNbO4 type, with
M = Y, La, Pr, Nd. The NbO4 tetrahedra in MNbO4 are
strongly distorted with two types of NbO distances equal
[218]. Neighboring NbO4 tetrahedra
to1.9522 and 1.8359 A
are placed in close proximity, with distances from niobium
. For these
to neighboring oxygen atoms of about 2.42.5 A
reasons, NbOx polyhedra in MNbO4 are often considered
as intermediate between isolated NbO4 tetrahedra and
edge-sharing chains of NbO6 octahedra with two long
NbO bonds [218220].
93
Nb NMR spectra of MNbO4 dier signicantly from
the spectra of compounds with well-dened NbO6 octahedra in their structure. Thus the 93Nb NMR static spectrum
of YNbO4 is very broad, and at low magnetic elds it was
practically impossible to record this spectrum without distortions (Fig. 34A) [54]. Computer simulation of this spectrum has produced a very large quadrupolar coupling
constant of 80 MHz and the isotropic shift unusually
shifted to 800 ppm. Similar parameters were also found
for LaNbO4 (CQ = 70 MHz, diso = 650 ppm) (Table 7).
Ultrahigh eld static NMR experiments (Fig. 34B) signicantly improve the situation: it was possible to obtain
a complete set of the chemical shielding and quadrupolar
tensors parameters which describe precisely the static spectra obtained at two elds (Fig. 34 and Table 7).
Crystalline structures of rare earth niobates MNbO4 are
very similar. Unfortunately, almost all these compounds
are paramagnetic, which complicates their 93Nb NMR
studies. When paramagnetic eects are not as strong, e.g.
for Pr and Nd, it was possible to obtain 93Nb NMR spectra
and to measure the corresponding NMR parameters
(Table 7). The 93Nb isotropic chemical shifts in PrNbO4
and NdNbO4 dier considerably from YNbO4 and
LaNbO4, most likely due to paramagnetic eects.
Thus, the 93Nb isotropic chemical shifts for four-coordinated Nb sites occur at the lower eld range (400 to
900 ppm) compared with the six-coordinated Nb sites
(900 to 1300 ppm). Due to signicant distortions of

166

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

A
1

4
2

3
-600

10000

-5500
-900

-1200

5000

-6000

-1500

-5000

-10000

-15000

, ppm

1, ppm

1, ppm

2, ppm

2, ppm

E
1, ppm

1, ppm

-1200

-1100

-1000

-1000

rf

-1100

-1200

2, ppm

Fig. 31. Experimental and calculated 93Nb NMR spectra of LiNbO3 measured at 9.4 T. (A) 30 kHz MAS spectrum. Details of the central transition
+1/2 M 1/2 and two satellite transitions, 3/2 M 1/2 and 5/2 M 3/2, are shown as insets. Simulated spectra are shown below with parameters:
CQ = 22 MHz, gQ = 0.2, diso = 996 ppm, mr = 30 kHz. (B) 93Nb DAS NMR spectrum. (C) 93Nb 3QMAS NMR spectrum. (D) Pure phase 2D nutation
93
Nb NMR spectrum. (E) 93Nb STMAS spectrum. Spectra (BD) are reproduced with permission from Ref. [50]. Spectrum (E) is reproduced with
permission from Ref. [66].

the local environment in four-coordinated sites and their


pseudo-octahedral type, the quadrupolar coupling constants are usually very large (>70 MHz). The magnetic
shielding anisotropy is less than 200 ppm.

6.1.3. Five-coordinated compounds


Among the compounds studied so far, only two have
niobium in coordination state ve, Na5NbO5 and
CaNb2O6.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

167

A
B

1
1
2

-500

-1000
, ppm

-1500

-500

-1000

-1500

, ppm
D
C
-4
1

0
4
8

2
12
-800

-1000

160

-1200

80

, ppm

0
ppm

-80

-160

Fig. 32. Experimental and calculated 93Nb NMR spectra of BiNbO4. (A) 30 kHz MAS spectrum obtained at 9.4 T. (B) Static spectrum obtained at 21.1 T.
(C) 30 kHz MAS spectrum obtained at 21.1 T. Calculated spectra shown below experimental spectra in (A), (B) and (C) were calculated with the same set
of NMR parameters, CQ = 23 MHz, gQ = 0.35, diso = 963 ppm, dr = 150 ppm. (D) DQ STMAS spectrum recorded at 21.1 T, mr = 20 kHz.

O2
O1

O2

O1

O1

O2
O2

O3
O1

Nb

La

O1
O1
O3

2000

1000

-1000

-2000

-3000

-4000

-5000

, ppm
Fig. 33. 93Nb MAS NMR spectra of La3NbO7 obtained at 9.4 T (35 kHz). (1) Experimental spectrum. (2) Simulated spectrum with the following
parameters: CQ = 49 MHz, gQ = 0.21, diso = 980 ppm, gd = 0, dr = 0 ppm, mr = 35 kHz. The structure of La3NbO7 consists of zig-zag chains of NbO6
octahedra with lanthanum ions occupying two types of polyhedra as illustrated.

The 93Nb NMR spectrum for Na5NbO5 is dierent


from the spectra of similar four-coordinated Nb compounds by having a much smaller quadrupolar coupling
constant (11 MHz) and a signicant high-eld shift of

diso. The complete set of NMR parameters for this compound was determined from NMR experiments at dierent magnetic elds (Fig. 35 and Table 7). The CS tensor
parameters and the tensors orientation were obtained

168

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

only from high-eld measurements. The small value of


the quadrupolar coupling constant reects the symmetric
nature of the Nb sites and their isolated character. The
high-eld shift as compared with four-coordinated compounds is due to the higher coordination number.
The isotropic 93Nb NMR shift of 990 ppm found for
CaNb2O6 is typical for ve-coordinated sites, and the relatively large CQ  50 MHz is the result of orthorhombic
symmetry.
Therefore, the following 93Nb NMR features have been
revealed for ve-coordinated Nb sites. In comparison with
four- and six-coordinated sites, ve-coordinated Nb demonstrates an intermediate range of the isotropic chemical
shifts. The quadrupolar coupling constant depends on the
local symmetry. It is small for symmetric isolated
sites, and large for pseudo-octahedral sites with one long
NbO bond.

6.1.4. Seven- and eight-coordinated compounds


Two of three studied 7- and 8-coordinated Nb compounds, LaNb5O14 and NaNb3O8, have several non-equivalent Nb sites in their crystal structure. Only in Cs3NbO8 is
there a single eight-coordinated site according to the
reported crystal structure.
The structure of LaNb5O14 consists of three types of
NbOx polyhedra, edge-sharing pentagonal NbO7 bipyramids forming chains, which are interconnected by cornersharing NbO6 octahedra. The 93Nb NMR spectrum for this
compound is very complicated due to presence of three different Nb sites with overlapping resonances. To obtain reliable 93Nb NMR parameters for all three sites the high-eld
DQ STMAS technique was applied. The 93Nb DQ STMAS
spectrum of LaNb5O14 has isotropic shifts at 1200,
1230, and 1267 ppm, corresponding to two octahedral
NbO6 and one NbO7 polyhedron, respectively [55].

2000

1000

-1000

-2000

-3000

-4000

-5000

-1000

, ppm

-2000

, ppm

Fig. 34. 93Nb NMR static spectra of YNbO4 obtained with a quadrupolar solid-echo pulse sequence at (A) 9.4 T and (B) 21.1 T. The experimental spectra
are show at the top. The spectra at the bottom were calculated with the following parameters: CQ = 82 MHz, gQ = 0.38, diso = 845 ppm, gr = 0.48,
dr = 200 ppm, a = 165, b = 16, c = 100.

-2700 -3000

-3300

2
-700

-800

-900

, ppm
93

-1000

-1100

4000

2000

-2000 -4000 -6000

, ppm

Fig. 35. Nb MAS NMR spectra of Na5NbO5 obtained at (A) 21.1 T and (B) 9.4 T. The experimental spectra are show at the top. The spectra below were
calculated with the following parameters: (A2) CQ = 11.1 MHz, gQ = 0.01, diso = 903 ppm, gr = 0.6, dr = 120 ppm, mr = 8 kHz. (B2) CQ = 11.1 MHz,
gQ = 0.01, diso = 903 ppm, gr = 0, dr = 0 ppm, mr = 30 kHz. Also shown on the right are ne details in the satellite transitions 3/2 M 5/2 and
3/2 M 1/2, and the central transition +1/2 M 1/2 together with simulations. The structure of Na5NbO5 is composed of isolated sodium and niobium
square-pyramids as illustrated on the left side.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

NaNb3O8 forms a channel structure comprised of two


types of chain-forming niobium polyhedra, edge-sharing
NbO8 dodecahedra and edge-sharing distorted pentagonal
bipyramids NbO7 [221]. However, for this compound,
93
Nb NMR experiments have been performed so far only
at 9.4 T, where it was very dicult to determine accurate
93
Nb NMR parameters for two non-equivalent Nb sites.
The 93Nb NMR static spectra were too broad to distinguish the two individual components with typical secondorder features of the central transitions. Nevertheless, it
was clear, that these spectra were indeed a superposition
of at least two sub-spectra, both strongly shifted up-eld.
From the high-speed 93Nb MAS spectrum the isotropic
shifts for both Nb sites were estimated. For seven-coordinated niobium sites the isotropic shift was found at ca.
1250 ppm and for eight-coordinated niobium sites at
1500 ppm.
According to the crystal structure, there is only one
type of eight-coordinated Nb site in the Cs3NbO8 compound, the only Nb site in the structure. Yet, the 93Nb
NMR experiments indicate the presence of at least two
overlapping components. Despite fairly symmetric oxygen coordination, four of eight ONb bonds are at
, and other four at 2.05 A
. The 93Nb NMR spec1.99 A
tra show quite strong quadrupolar coupling interactions,
CQ = 62 MHz, large CSA, and a considerable up-eld
shift (1600 ppm).
When compared with six-coordinated niobium compounds, seven- and eight-coordinated Nb sites usually
demonstrate high-eld shifts, while the magnitude of the
quadrupolar coupling constant is mostly determined by
the local symmetry.
CSA and quadrupolar parameters for a large number of
individual niobium-containing oxide materials are summarized in Table 7. Many of these experimental results are
presented here for the rst time.
6.2.

93

Nb NMR chemical shift scale

By analyzing the data reported in Table 7 it is straightforward to make several important conclusions regarding
the relationship between the isotropic 93Nb NMR chemical
shifts and the niobium coordination environment.
For four-coordinated Nb sites, the isotropic chemical
shifts corrected for the second-order quadrupolar perturbations occur from 650 to 950 ppm, with
CQ > 70 MHz.
For ve-coordinated Nb sites the isotropic chemical
shifts are observed in the range from 920 to 990 ppm,
CQ  1050 MHz, dr  200 ppm.
For six-coordinated Nb sites the isotropic shift diso varies from 900 to 1300 ppm, CQ from 1 to 100 MHz, and
dr from 0 to 300 ppm. The range of 900 to 1000 ppm
for diso is typical for cubic symmetry. For six-coordinated
Nb sites with non-cubic symmetry, the 93Nb isotropic
chemical shifts are inuenced by the ionic character of
the niobium sub-lattice. In niobates of M(+1, +2, +3) ele-

Cubic M(+1,2,3)
-900

M(+4)

169

M(+5)

-1000 -1100 -1200 -1300 -1400

NbO8
NbO7
NbO6
NbO5
NbO4

-500

Fig. 36.

-1000

-1500

iso, ppm

-2000

93

Nb NMR chemical shift scale for NbOx polyhedra.

ments the niobium sub-lattice has more anionic character,


and the isotropic shifts are in the range from 1000 to
1100 ppm. When M(+4) or M(+5) is present, the niobium sub-lattice has more cationic character and the isotropic chemical shifts are from 1200 to 1300 ppm. The
value of the quadrupolar coupling constant for six-coordinated Nb sites varies over a wide range and depends on the
site symmetry. The chemical shift anisotropy is large only
when atoms other than oxygen are present in the rst coordination sphere, for example, uorine.
For seven-coordinated niobium sites the isotropic shift
varies from 1200 to 1600 ppm while for eight-coordinated niobium sites the isotropic shift occurs at elds
higher than 1500 ppm.
On this basis the 93Nb NMR chemical shift scale has
been proposed for niobium compounds (Fig. 36) [55].
7.

181

Ta NMR data compilation

Only a very limited number of solid-state 181Ta NMR


studies has been reported to date, including several early
works on K[TaO3] [16,57,62,63], MI[TaY4] (MI = Cu, Tl;
Y = S, Se, Te) [58], tantalum metal [59], and TaY2
(Y = S, Se) [60,61]. The 181Ta chemical shift range in solution extends over 3450 ppm, and is limited by [TaCl6] species at the low-eld limit and by [Ta(CO)6] at the higheld limit [56]. The shielding sensitivity of 181Ta is about
1.6 times that of 93Nb and is comparable with that of
95
Mo. The 181Ta NMR spectrum of cubic KTaO3 consist
of a rather narrow (5 kHz) central transition, with satellite
transitions spread out over 100 kHz around the central
line. Doping KTaO3 with niobium, lithium and sodium
[63] leads to considerable line broadening due to distortion
of the cubic structure in pure KTaO3.
8. Paramagnetic eects in
spectra

51

V and

93

Nb solid-state NMR

In general, the eects of paramagnetic metal ions on


NMR spectra (chemical shifts, relaxation times, etc.) are

170

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

well described for paramagnetic molecules and paramagnetic solids with low and intermediate concentration of
paramagnetic centers [223225]. Paramagnetic shifts are
particularly useful in discovering medium and long-range
structural information. We will not discuss here the origin
of all the dierent paramagnetic contributions to the chemical shift referring our readers to the cited literature. We
note, however, that all the paramagnetic eects under consideration may be eectively incorporated in the secondrank hyperne interaction (HFI) tensor, which can be
directly probed by magnetic resonance experiments. The
Fermi contact and pseudocontact contributions to the
chemical shift are the most important in the context of
the present discussion.
The contact shift is caused by the contact term in the
spin-Hamiltonian as was rst derived by Fermi. This shift
is associated with the isotropic contribution to the HFI tensor from the electron spin density located directly at the
nucleus site. For paramagnetic ions of the VB Group,
and especially for vanadium, unpaired electrons are not
completely concentrated at the ion location, but are also
transferred via direct delocalization and spin polarization
to remote atoms of the surrounding ligands. The pattern
of electron spin delocalization is not easily predictable from
rst principles, especially for solids. This is one of the reasons why the information obtained from contact shift values is often limited to the general picture of the electron
spin delocalization over the nearest nuclei active in NMR.
In disordered paramagnetic systems with magnetically
anisotropic paramagnetic centers there is an additional orientationally averaged contribution to the isotropic value of
the chemical shift called the pseudocontact, or dipolar,
shift. It originates from the anisotropy of the net electron
spin moment connected with a susceptibility tensor xed
at the molecular structure. The most common representation for a pseudocontact shift is given by


1
3
2
2
dpc
Dv
Dv
3
cos
h

1

sin
h
cos
2u
9
ax
12pr3
2 rh
with
Dvax vZZ 

vXX vYY
;
2

vrh vXX  vYY

10

where Dvax and Dvrh are the axial and rhombic anisotropy
parameters of the magnetic susceptibility tensor v of a
paramagnetic metal, while two angles h and / dene the
orientation of an electronnucleus vector in the frame of
the magnetic susceptibility tensor. Somewhat dierent
expressions for the pseudocontact shift can be found in
[225]. Interpretation of the pseudocontact shifts depends
to a lesser extent on the electronic structure theory of
molecules or solids, comparing with similar interpretations of the Fermi contact shifts. Nevertheless, such interpretation often requires comparison to a diamagnetic
analog with similar geometry and charge distribution.
For example, for lanthanides unpaired electrons are
mainly located in inner orbitals with little or no overlap

with orbitals of the surrounding ligands. This makes the


Fermi contact shifts negligible compared with the pseudocontact contributions to the paramagnetic shift. In some
cases pseudocontact shifts can be experimentally mea from the parasured even for nuclei located as far as 40 A
magnetic center.
Therefore, by measuring local elds at the location of a
nucleus, it is possible to obtain information on the spindensity transfer from the paramagnetic ion, and also to
estimate the distance between the observed nucleus and
the paramagnetic ion.
This simplied picture of electron-nuclear interactions
becomes incomplete or even incorrect for solids with a high
concentration of paramagnetic centers and especially for
magnetically ordered materials. In these situations other
approaches originating from solid-state physics need to
be applied [226230].
While V5+ containing phases are directly characterized
by conventional 51V NMR, this is not the case for materials with vanadium atoms in lower oxidation states. It is
possible, however, though indirectly, to obtain information concerning the nature, location and the oxidation
state of vanadium centers from NMR spectra of neighboring atoms. An example of such an investigation for
vanadiumphosphorus systems was reported by Lashier
et al. [231] and later by Tuel et al. [232]. They have
applied a so-called 31P spinecho mapping technique to
probe vanadium paramagnetic centers by observing the
31
P chemical shift over a very large spectral region. In
compounds containing only diamagnetic V5+ centers the
31
P NMR chemical shifts are in the range 2040 ppm; at
the same time for compounds containing paramagnetic
V4+ and V3+ the 31P NMR chemical shifts range from
1600 to 2600 ppm and at around 4650 ppm, respectively.
The 31P NMR line shift is directly proportional to the
density of unpaired electrons. Thus, the 31P NMR chemical shift has provided information about the number of
paramagnetic vanadium species in the rst coordination
sphere of phosphorus atoms, as well as on the oxidation
state of such species.
Applying the spinecho mapping technique to 51V under
ultra-high speed MAS (and large spectral width) conditions, it is possible to detect 51V NMR signals of V5+ atoms
in close proximity to paramagnetic centers.
In this work we report our recent 51V NMR data for a
number of rare earth vanadates where paramagnetic eects
are caused by the presence of paramagnetic cations. We
also summarize results available for vanadium bronzes,
where some vanadium is in a paramagnetic V4+ state,
and for Cuban-like vanadium compounds, where closely
spaced paramagnetic centers form diamagnetic pairs when
placed in persistent magnetic elds.
Since all these compounds are true paramagnetics, special
care should be taken in choosing appropriate experimental
conditions while performing NMR experiments, i.e. the
amount of the sample in the MAS rotor should be as small
as possible.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

8.1. Presence of paramagnetic cations


Early studies of the eects of the paramagnetic centers
on the 51V NMR spectra were usually performed with
CW-NMR spectroscopy. The most important results of
51
V CW-NMR in vanadium oxides, bronzes, and various
vanadates of s-, d-, and f-elements have been summarized
by Pletnev et al. [22], together with some experimental
spectra, and several examples of the cluster MO-calculations of the electric eld gradients and the magnetic
interactions.
Examples of 51V MAS NMR spectra for several vanadates of rare earth and 3d-elements are presented in
Fig. 37, with the corresponding spectral parameters given
in Table 8. The 51V quadrupolar coupling constant found
for iron orthovanadate FeVO4 is rather small, while for
rare earth orthovanadates the quadrupolar coupling constant exceeds those found in meta- and orthovanadates of
the Group I and Group II elements. The asymmetry
parameter of the quadrupolar tensor is zero for most of
these compounds. It was reported earlier by Pletnev et al.

171

[233,234] that the 51V NMR isotropic chemical shifts for


Cr, Fe, Co, and Ni orthovanadates are shifted to low eld
compared with that of the reference KVO3 solution. Indeed
this is the case for FeVO4, where the isotropic chemical
shift in the 35 kHz MAS spectrum is at 17,000 ppm, i.e.
at much lower eld to the reference sample. Because of
between vanadium
the considerable distance of 3.4 A
atoms and the paramagnetic centers, the direct contact
interactions of 51V nuclei with electrons located on the partially occupied 3d shells is unlikely. This strong eect on
the isotropic chemical shift can be explained by transfer
of electron density from the magnetic d-electrons via the
atomic orbitals of nonmagnetic ions, vanadium, and
oxygen.
The isotropic 51V NMR chemical shifts in rare earth
orthovanadates can be formally presented as:
diso d0 dp ;

11

where d0 is the chemical shift originating from the electronic structure of the lled orbitals, and dp is the shift

*
4

20000

18000

16000

14000

, ppm

*
3

*
2

*
1

4000

2000

-2000

-4000

, ppm
Fig. 37. 51V MAS NMR spectra of MVO4 vanadates with paramagnetic cations: (1) PrVO4, (2) YbVO4, (3) EuVO4, and (4) FeVO4. The isotropic lines are
marked with asterisks. Note an extremely large shift of +17,000 ppm observed in the Fe-containing sample. The spectra were obtained at 9.4 T using a
2.5 mm MAS probe with mr = 30 kHz.

172

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

Table 8
51
V NMR parameters for several vanadates of rare earth and 3d-elements
Compound
FeVO4
CeVO4
PrVO4
NdVO4
EuVO4
HoVO4
ErVO4
TmVO4
YbVO4
a
b

CQ (MHz)a
1.7
3.5
5.59
5.5
4.95
4.55
4.51
4.43
4.25

gQa

0.9
0
0
0
0
0
0
0

diso (ppm)b
17,000
559
253
242
935
0
30
90
100

dr (ppm)b

200
520
550

Ion conguration
5

3d
4f1
4f2
4f3
4f6
4f10
4f11
4f12
4f13

2S+1

LJ base state

11

S5/2 (11)
F5/2 (6)
3
H4 (9)
4
I9/2 (10)
7
F0 (1)
5
I8 (17)
4
I15/2 (16)
3
H6 (13)
2
F7/2 (8)
2

Values of CQ and gQ are from Ref. [22].


Values of diso and dr are obtained in this work.

from interactions of the vanadium atom with unpaired


electrons of the M3+ ions.
When diso = d0, as found in diamagnetic vanadates,
the isotropic chemical shifts are in the range typical
for vanadium in an oxygen environment, i.e. as in
CeVO4, LuVO4, ScVO4, YVO4, and LaVO4 (Table 8).
This conrms that the d0 values are about the same
for the whole range of the rare earth orthovanadate.
Therefore, all the changes in diso observed for rare earth
vanadates are due to changes in dp. Also, there was no
correlation found between the values of the isotropic
chemical shifts and the g-factor anisotropy in ESR spectra [235,236]. Therefore, the contribution of the pseudocontact interactions to the chemical shift can be dismissed as negligible.
Vanadates containing mixed paramagnetic and diamagnetic cations are of interest for developing advanced
lithium ion batteries. One such system based on LiCoxNi1xVO4 compounds with paramagnetic Co and Ni
was recently studied by Stallworth et al. [192]. In their
preparations they varied the Co/Ni ratio, with x assuming
values from 0 (only Ni present), 0.2, 0.5, 0.8, and 1.0
(only Co present). The workers reported observing two
resonances in the 51V NMR spectra. One resonance in
the range 2500 to 4200 ppm having a Gaussian line shape
and strongly aected by paramagnetic interactions was
attributed to tetrahedral vanadium sites. The second resonance with the isotropic chemical shift of 630 ppm and
exhibiting rst-order quadrupolar broadened satellites was
assigned to octahedral sites. This contradicts, however,
the correlation presented above in Section 5. The
latter line at 630 ppm exhibiting the rst-order quadrupolar interactions should instead have been attributed
to tetrahedral vanadium sites. Moreover, diamagnetic
orthovanadates LiCdVO4, LiZnVO4, and LiMgVO4
where vanadium has a tetrahedral oxygen environment
shows very similar 51V NMR spectra. It is clear that these
tetrahedral sites are surrounded by diamagnetic Li and
not aected by paramagnetic interactions. At the same
time, the resonance at 25004200 should correspond to
octahedral sites and tetrahedral sites with paramagnetic
neighbors.

8.2. Systems with vanadium in mixed oxidation states


Vanadium bronzes, compounds somewhat similar to
metals in their properties, are formed by incorporating M
elements into the vacancies of the VOn oxide structure.
The charge balance is achieved by transferring two valence
electrons of M ions to vanadium ions, which are nominally
present in two oxidation states. To underline the fact that
vanadium bronzes are actually oxides, sometime they are
called oxide bronzes, or non-stoichiometric compounds
incorporating one-, two-, and three-charged ions. Element
M incorporated into the crystal structure of the oxide normally occupies only a fraction of the available voids, with
the lling factor depending on the ionic radius of Mn+.
By varying the concentration of M, it is possible to obtain
a number of MxVOn phases with dierent composition
[237]. Thus for V2O5 there are four such phases, a, b, c,
and d. When x is not exceeding 0.04, the a-MxV2O5 phase
is formed, representing solid solutions of M1+,2+,3+ ions
incorporated into V2O5 [238]. The structure of this a-phase
is similar to V2O5. The b-phase is formed when
0.2 < x < 0.4. In the b-phase there are three non-equivalent
vanadium sites, two of them represent deformed octahedral
coordination, and the third site is ve-coordinated. Further
incorporation of Mn+ leads to formation of c and d phases
at x greater than 0.6.
One may expect that dierent types of phase transitions
in MxV2O5 obtained upon varying x could be followed
with 51V NMR. However, while studying LixV2O5
(0.4 < x < 1.4) Nakamura et al. [239] observed that for different values of x the isotropic 51V shifts did not change
signicantly. Based on this observation it was concluded
that the VO5 square pyramidal structure of V2O5 is
retained without drastic structural modications.
As already discussed above, the isotropic 51V shift is not
a very informative probe of the vanadium environment, the
most informative being the chemical shift anisotropy and
the quadrupolar coupling constant. Indeed close inspection
of the 51V NMR spectra presented in the above mentioned
paper shows that variations in x do lead to changes in
vanadium environment. At loading x = 0.4 the structure
is similar to V2O5, with one of the sites being of the b-phase

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

or a residual part of the a-phase. The bronze of b or c type


is formed at a value of x of about 0.8, as indicated by a
broad anisotropic line with a chemical shift of 797 ppm.
The isolated tetrahedral sites are formed at even higher x
loadings, x > 1, which is supported by the small anisotropy
observed and the moderate value of the quadrupolar constant. This type of transformation seems to be typical for
other MxV2O5 bronzes, where several consecutive phase
transitions can be observed up to x = 0.8: a b b 0 c.
Whereas at x > 0.8 the bronze structure is usually breaking
apart.
The b(b 0 ) type vanadium bronzes of Li, Na, K, Ag, Bi,
and Sr are structurally similar (Table 9). At normal conditions these compounds are paramagnetic metals with wellpronounced conductivity along the b axis. At lower temperature these compounds transform into non-metallic
phases, and at the liquid He temperature convert either
into ordered ferromagnetic phases as in Na0.33V2O5, or into
a diamagnetic phase as in Ca0.3V2O5. The crystal structure
of these bronzes can best be described as a strongly distorted V2O5 phase with three non-equivalent vanadium
sites, V1, V2, and V3. It is often assumed, although it is
not certain, that the paramagnetic V4+ ions, or unpaired
Table 9
Parameters of the

173

electrons, are located either in V1 or V2, or in both, while


V5+ ions occupy only V3. From the data presented in Table
9, it is clear that the line at 900 ppm is typical for all
bronzes of the b type, and perhaps can be attributed to
V5+ located near V4+. If, however, V4+ ions or unpaired
electrons occupy V1 and V2 sites, than V5+ shifted to the
high eld may also occupy V1 and V2 sites. In Ca0.3V2O5
and Cs0.35V2O5 bronzes 51V NMR signals were detected
at 2120 and 1565 ppm respectively, which is very close to
signals from diamagnetic V4+OV4+ pairs in b-VO2. This
correlates with the non-magnetic state of Ca0.3V2O5 at
liquid He temperature. It is interesting, that in Ca0.3V2O5
the 51V NMR signal is at 4766 ppm, very close to metallic
a-VO2. In similar metallic bronzes 51V NMR signals can be
observed from 1718 to 739 ppm. The more metallic the
bronze becomes, i.e. moving from Na0.33V2O5 to
Na0.67V2O5, the more the 51V NMR signal is shifted
towards the signal found in pure vanadium metal.
Some representative 51V NMR spectra for vanadium
bronzes and VO2 are shown in Figs. 38 and 39. It is important to recognize, that the interpretation of such spectra is
not always straightforward, which sometimes leads to
heated debates and controversy in the literature [240].

51

V NMR spectra in oxide bronzes and compounds containing V4+V4+

Compounds

CQ (kHz)

gQ

dr (ppm)

gr

diso (ppm)

Ref.

Cuban-like compounds
V2S4(S2COEt)4

3120

0.4

960

0.001

This work

V2S4(S2COi-Pr)4

3120

0.4

960

0.001

V2S4(S2CNBu2)4
4,4 0 -Bipyridinium cations, (C10H10N2)[(VO2)4(PO4)2]

3126

0.40

956

0.001

70
54
105
82
56
135.0
580.1
593.4
599.2

5200

This work

6940
1900

0.46
0.4

1151
175

0.88

2113
4788

[242]
[242]

2100
4000

0.9
0.9

270
1000

1.0

This work

0.9

1000
800
800
370
835
1600

0.8
0.92

2120
4766
1447
1565
775
810
739
869
1200
900
620
900
435
900
1718
1381

Metal
metal V
Oxides
b-VO2
a-VO2
Bronzes
Ca0.3V2O5
Bi4V2O10.6
Cs0.35V2O5

Na0.67V2O5
Li0.3V2O5
Na0.33V2O5

4000

K0.25V2O5
Ag0.35V2O5
Sr0.17V2O5
Ba0.17V2O5
Composites
PANI-V2O5-nanocomposites

860
1250

0.01

8500

This work

This work
[241]

[189]
This work

This work
This work
This work
This work
This work
This work
This work
[243]

174

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

For example, the high-temperature phase of a-VO2 is


clearly metallic with characteristic anisotropy in properties,
and at the same time this phase is characterized by the
abnormal 51V NMR chemical shift at 4788 ppm, and
opposite in sign to the Knight shift in vanadium metal.

This behavior can be explained by dominating contributions of the electron polarization transfer between ground
electron orbitals and those with unpaired d-electrons. In
a similar fashion, to describe the non-metallic diamagnetic
electronic state of the low-temperature b-VO2 phase, it was

*
2

4000

2000

-2000

-4000

-6000

-8000

, ppm
Fig. 38. 51V NMR MAS spectra of polycrystalline vanadium bronzes recorded at 9.4 T under 30 kHz MAS, (1) Ba0.17V2O5 and (2) Ca0.3V2O5. Isotropic
lines are marked with asterisks.

*
2

*
1

6000

4000

2000

-2000

-4000

, ppm
Fig. 39. 51V MAS NMR spectra of two vanadium compounds with V4+V4+ pairs in the structure. (1) VO2 and (2) V2S4(n-Bu2NCS2)4. Isotropic lines are
marked with asterisks. The structure of a [V2S4]4+ unit is shown above. Due to a very short distance between two V4+ ions these compounds are
diamagnetic as also supported by 51V NMR. The spectra were obtained at 9.4 T using a 2.5 mm MAS probe at mr = 30 kHz.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

175

suggested to use either a model involving mixed


oxidation states, V3+ and V5+, or a model of the diamagnetic V4+OV4+ (S = 0) pairs. The latter model is more
prevalent, and is supported by the isotropic 51V NMR shift
values found in four model compounds with Cuban-like
structures containing diamagnetic pairs V4+SV4+
(Fig. 39).
From the practical point of view 51V CW-NMR spectroscopy used in early studies on vanadium bronzes
[233,234,237,240] has certain advantages over modern Fourier-transform NMR spectroscopy. The most important is
a possibility of recording extremely broad spectra with
almost unlimited spectral width. At the same time, pulsed
NMR experiments with the same samples would require
extremely broad sweep widths, short high-power RF pulses
and very high spinning speeds if MAS is to be applied. On
the other hand, modern NMR approaches bring increased
spectral resolution and the ability to resolve non-equivalent
vanadium sites, which is impossible by CW-NMR.
9. Recent applications of solid-state
oxide materials
9.1. Applications of solid-state

51

51

V and

93

Nb NMR in

V NMR

The last two decades have seen a dramatic increase in


the number of 51V NMR applications to a variety of solids
ranging from materials science and biology to solid-state
physics. In heterogeneous catalysis, solid-state 51V NMR
has become almost a routine characterization technique.
The seeming simplicity of 51V NMR has brought about a
very wide user base, yet has had a somewhat negative
impact as well. The main problem remains with inadequate
NMR hardware which is being applied to study very challenging and complex systems. Herein, we have selected a
few of the most representative examples of successful 51V
NMR applications from the recent literature, covering
major research areas, including bio-structural chemistry,
materials science, and catalysis.
9.1.1. Bio-structural chemistry
Perhaps the most impressive example of the solid-state
51
V NMR spectroscopy applied directly to probe vanadium
centers in proteins has been reported very recently by Pooransingh-Margolis and coworkers while studying 67.5-kDa
vanadium chloroperoxidase [136]. Each molecule of this
protein contains a single vanadium atom, which translates
into only 1 lmol of vanadium spins in the entire sample.
Despite this very low concentration of vanadium sites,
the authors were able to detect and analyze the spinning
sidebands of the central and satellite transitions (Fig. 40).
The quadrupolar and chemical shift anisotropy tensors
were determined by numerical simulations of the spinning
sidebands and the line shapes of the individual spinning
sidebands of the central transition. For these vanadium
sites the authors reported a value for the quadrupolar coupling constant CQ of 10.5 MHz and the chemical shift

Fig. 40. 51V solid-state NMR spectra of vanadium chloroperoxidase


acquired at 14.1 T at two MAS spinning speeds of 17 kHz (1) and 15 kHz
(2). The isotropic line at 520 ppm is marked with an asterisk. The inset
shows spinning sidebands from the satellite transitions indicating large
CQ. Each spectrum took about 5 days to acquire. Reproduced with
permission from Ref. [136].

anisotropy dr of 520 ppm. DFT calculations of the NMR


spectroscopic observables for an extensive series of active
site models indicates that the vanadate cofactor in these
enzymes is most likely anionic with one axial hydroxogroup and an equatorial plane consisting of one hydroxoand two oxo-groups. This approach has yielded the
detailed coordination environment of the metal centers,
which is unavailable from other experimental measurements. Solid-state 51V NMR is expected to be generally
applicable for studies of diamagnetic vanadium sites in
other metalloproteins and similar systems.
9.1.2. Materials chemistry
Hybrid organicinorganic materials based on vanadium
oxide have received signicant attention due to their ionic
and electronic properties with potential applications in various technologies, such as reversible cathodes in lithium
batteries, electrochromic devices, or even in heterogeneous
catalysis. Many vanadium compounds exhibit a layered
structure allowing the reversible intercalation of a wide
range of organic molecules to occur.
Durupthy et al. have recently studied vanadium oxide
foams prepared by mixing V2O5 with hydrogen peroxide
and solution of 1-hexadecylamine [202]. Despite the disordered nature of these foams, using solid-state 51V NMR the
authors have been able to successfully identify a variety of
polyoxovanadates contained in them. It has been shown
that the local environment of vanadium as probed by 51V
MAS NMR consists of octahedral and tetrahedral vanadium sites corresponding to decavanadate [HxV10O28](6x)

176

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

and [V4O10]4/[V2O7]4 polyanions. However, the relative


amounts of dierent polyoxovanadate species could not
be controlled via synthetic means. The layered structure
of foams is most likely formed by a self-assembly of
amines, while [HxV10O28](6x) and [V4O10]4/[V2O7]4
polyanions intercalate into interlayer spaces. In a similar
fashion, the structure of the tailor-made macroporous
vanadium oxide foams has been examined by 51V NMR
in [244] (Table 5). It has been revealed that the local structure of vanadium sites in these foams is similar to that
reported earlier for V2O5 xerogels [148].
Grey et al. [245,246] used variable temperature 51V MAS
NMR to study the synthesis of Li1+rV3O8 (r = 0.10.2)
materials via a xerogel precursor route. Li1+rV3O8 materials are among several other layered lithium-containing
vanadates with attractive electrochemical properties for
applications in rechargeable lithium batteries. These materials generally consist of V3O8 layers also found in hewettite, CaV6O169H2O, with octahedrally and pentacoordinated vanadium atoms. Intercalation can occur
between the layers. During synthesis the solid component
of the xerogel undergoes a series of phase transitions
involving layered phases with decreasing water concentrations, while the hewettite framework is largely maintained
throughout the reaction. In the case of the dried liquid
xerogel component the formation of anhydrous Li1+rV3O8
occurs through decomposition of lithium vanadates, which
was followed rst by formation and then by progressive
dehydration of the hydrated hewettite structure.
Valence states of vanadium in VO2 have been studied
with 51V MAS NMR in [242] in order to reveal if VV pairs
consist of either V5+V3+ or V4+V4+. To answer this challenging question, the authors applied the high-speed MAS
technique with spinning speeds up to 35 kHz at two dierent magnetic elds, 9.4 and 14.1 T. These spectra unambiguously show the presence of a single V4+ site in the lowtemperature non-metallic b-form of VO2. Variable-temperature 51V MAS NMR studies performed in the temperature
range from 25 to 90 C indicate that the phase transition
from non-metallic b to metallic a- VO2 is also accompanied

by a dramatic change in the Knight shift value, from 2115


to 4788 ppm.
Michalaka and Mohammad were the rst to directly
detect 51V NMR spectra from more than two valence states
of vanadium present simultaneously in the same compound
[247]. They identied three individual resonances in the 51V
NMR spectra of the low-temperature phase of a-NaV2O5.
These three resonances were attributed to three dierent
vanadium valences in the system. NMR data were compared with the structural models developed for this
compound.
9.1.3. Catalysis
Heterogeneous vanadia-based catalysis has been a traditional stronghold for solid-state 51V NMR spectroscopy
starting with its earliest applications in late 1950s, almost
immediately after the discovery of the magnetic resonance
phenomenon. We have reviewed the subject of 51V NMR in
catalysis on several occasions, including our rst comprehensive review published in this very journal in 1992 [34].
Another review appeared in the Encyclopedia of Nuclear
Magnetic Resonance in 1996 [44], and the most recent
was published in 2003 in Catalysis Today [45]. A considerable number of research papers on this subject has been
published since then, many using solid-state 51V NMR as
their primary research tool. In this review, we will discuss
several of the many recent publications, focusing mostly
on those having methodological interest for development
of the 51V NMR technique. Some examples are selected
to specically illustrate recent NMR hardware developments, including the availability of extremely high spinning
speeds for MAS in excess of 35 kHz, as well as the wider
accessibility of ultrahigh magnetic elds for solid-state
NMR.
Binary catalysts containing vanadia supported on inert
supports have been studied in detail not only by NMR
but by many other techniques, and most recently by computer modeling. The model shown in Fig. 41 is one of several introduced to describe surface vanadia species and
adopted by many research groups.

Fig. 41. Model of the surface structure in VOx/TiO2 catalysts. Reproduced with permission from Ref. [6].

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

Although this particular model was developed for


VOx/TiO2 catalysts, it has a more general applicability to
other similar binary vanadia systems. Depending on the
preparation technique as well as on the surface concentration of vanadia, the support nature, and the treatment
conditions, several distinct types of VOx species can be
identied and characterized by solid-state NMR.
Isolated or associated surface vanadia species are formed
at low vanadia loadings, normally at loadings not exceeding a theoretical monolayer. The structure of these species
depends on the number of chemical bonds connecting
vanadium to the support, and also on the treatment conditions. This can be the best illustrated by samples prepared
via chemical grafting. The grafting technique suggests
direct chemical reaction between surface hydroxyl groups
of the support with vanadium compounds being deposited,
usually VOCl3 or VO(OR)3:

The next step of the catalyst preparation involves a


hydrationdehydration procedure at 350 C. Regardless
of the initial grafting temperature, the corresponding 51V
MAS NMR spectra always showed three similar groups
of signals (Fig. 42B, spectrum 2):V1 with diso = 537 ppm,
gr = 0.1, dr = 250 ppm, CQ = 2.1 MHz, gQ = from 0 to 1;
V2 with diso = 570 ppm, gr = 0.9, dr = 230 ppm,
CQ = 4.3 MHz, gQ = 0.7; V1 with diso = 645 ppm,
gr = 0.3, dr = 600 ppm, CQ = 7 MHz, gQ = 0.4.
Only MAS spinning of these samples at 35 kHz (at 9 T)
frees all isotropic lines from overlapping with MAS spinning sidebands. From their NMR parameters it is straightforward to assign V1 to trigonal VO4 pyramids, V2 to
associated tetrahedral sites of Q2 type, and V3 to distorted
octahedral sites (see above). High-eld isotropic shifts
found for V3 sites are most likely due to strong interactions
between vanadium and TiO2, for example, via two or three
VOTi bonds. At the same time, V1 sites should correspond to vanadium weakly bonded to TiO2, i.e. via only
one or two VOTi bonds. This assignment correlates well
with the electronegativity of the atoms in the second coordination sphere. Isotropic 51V chemical shifts progress
towards more negative values as the electronegativity of a
second metal atom decreases. Consequently, the higher
electron density around vanadium in V3 species can be
attributed to their stronger interaction with TiO2 than
those of V2 and V1 species [249]. Taking into account complementary Raman data [250,251] it is plausible to suggest
that V3 species are monomeric distorted VO6 octahedra
with one short V@O bond which are interacting strongly
with the support.
Each group of signals in Fig. 42B, spectrum 2 in turn
consists of two to three closely overlapping lines. This
can be explained by vanadium interacting with OH groups
located not only on the densest plane (0 0 1), but also on
cleavage planes other than (0 0 1). Formation of at least
six to nine non-/equivalent vanadium sites indicates that
there should be also association of vanadium species:

VOCl3 nHOTi ! VOCl3n OTin


here n could be 0, 1, 2, or 3. The resulting surface vanadium species, their structure and the number of chlorine
atoms left in this reaction can be identied in the 51V
MAS NMR spectra by their isotropic chemical shifts and
the type and value of the chemical shielding anisotropies
[248].
For example, gas-phase grafting of VOCl3 onto an anatase surface at room temperature leads to formation of tetrahedral VOCl2(OTi) species with each vanadium having
only a single VOTi bond with the surface and the
two chlorine atoms in the rst coordination sphere
(diso = 350 ppm, gr = 0.2, dr = 230 ppm). Depositing
VOCl3 at 110 C results in stronger bonding of vanadium
to the surface as VOCl(OTi)2 species, now with two
VOTi bonds and only one chlorine atom in the rst coordination sphere (diso = 433 ppm, gr = 0.4, dr = 220 ppm).
Even stronger bonding of vanadium occurs at 250 C,
when VO(OTi)3 species form (diso = 520 ppm, gr = 0.3,
dr = 310 ppm, CQ = 2.2 MHz, gQ = 0.3).

+H2O

-200

-1000

, ppm

V1

V2

V3

after
reaction

initial

177

-1800

-500

-600

-700

, ppm

Fig. 42. (A) Static and (B) 35 kHz MAS 51V NMR spectra recorded at 9.4 T for VOx/TiO2 catalysts prepared by grafting technique. A2, B2 initial
catalysts, A1 after H2O adsorption, B1 after catalytic reaction.

178

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191
350  C

H2 O

VOCl2 OTi ! VOOH2 OTi ! VOOHOTi2


350

C

! VOOTi3
350  C

H2 O

VOClOTi2 ! VOOHOTi2 ! VOOTi3


350  C

2VOOHOTi2 !OTi2 VOOVOOTi2


OTi2 VOOVOOTi2 VOOHOTi2
350  C

! VOOTi3 OVOOTiOVOOHOTi2

According to complementary 1H MAS NMR data


[34,248,252], both bridging and terminal OH groups are
involved in bonding VOx species, with on average two
VOTi bonds found after the hydrationdehydration
procedure.
During catalytic reactions both the structure and content of surface sites undergo certain transformations. Thus
under DeNOx conditions the weakly bound V1 sites on the
TiO2 surface change signicantly, with their relative content increasing, and the sites themselves becoming more
symmetric. At the same time, V2 sites are practically unaffected under these conditions. Similarly, only minor
changes can be detected for V3 sites (Fig. 42B, spectrum 1).
Grafting VOCl3 onto a silica surface with subsequent
hydration and dehydration at T P 350 C almost always
leads to strongly bound VO(OSi)3 species having three
VOSi bonds with the surface. 51V NMR spectra of
dehydrated VOx/SiO2 samples prepared this way are characterized by an axially symmetric anisotropy of the CSA
tensor, with parameters typical for trigonal pyramids, i.e.

diso = 710 ppm, d^ = 460 ppm, gr = 0.04, dr = 480 ppm,


and CQ = 2.5 MHz (Fig. 43).
Upon hydration the surface vanadia species interact
readily with water molecules forming various surface complexes, i.e. VO3(OH), (VO3)n, VO2(OH)2, V10O26(OH)2,
VO(OH)3, V10O27(OH), (V2O7)4, or V10 O28 6 , depending
on the net surface pH, as was demonstrated using Raman
spectroscopy by Deo and Wachs [253]. This agrees with
recent 51V NMR data, which have also shown that surface
complexes remain attached to the surface, though not as
strongly as in the dehydrated state [148,254,255]. Surface
vanadia species in hydrated VOx/SiO2 are similar to those
found in V2O5nH2O gels.
Based on 51V and 17O NMR data, a total of ve dierent
vanadium species were identied and quantied in vanadia
gels at various stages of hydration [148,254,255]. Comparative analysis of 17O MAS and 3QMAS NMR, 51V MAS
NMR, and thermogravimetric data has provided an additional insight into the coordination states of water molecules during hydration, and led Fontenot et al. to
propose a model for the gel structure. Upon rehydration
of the layered gel there is observed to be a single preferred
site for initial water readsorption. Oxygen atoms of these
readsorbed water molecules are readily exchangeable with
all other types of oxygen sites even at room temperature
(Fig. 43).
The surface vanadia species described above have been
found not only in the catalysts prepared by grafting, but
also in vanadia-based catalysts prepared by other techniques: impregnation, mechano-chemical activation, co-

Fig. 43. Wide-line 51V NMR spectra recorded at 9.4 T for VOx/SiO2 catalysts prepared by grafting technique. (1) Initial catalyst after dehydration. (2)
Catalyst after H2O adsorption. Structures of vanadium sites in the dehydrated catalyst and in the catalyst after H2O adsorption are adopted with
permission from Ref. [255].

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

precipitation, or spray-drying. However, in all these catalysts other types of vanadia are also present in addition
to surface species, including larger polyanions, strongly
bound vanadium (SBV), binary phases, and V2O5.
Strongly bound vanadium (SBV) is often found in catalysts prepared via spray-drying of the mixture of TiO2
and vanadyl oxalate followed by a thermal treatment.
Under these conditions a coherent interfacial boundary is
formed between two crystalline phases, TiO2 and V2O5.
According to HREM and EDAX data this interfacial
boundary consists of closely spaced vanadium and titanium atoms in a 1:1 ratio forming a mixed-cation layer.
The structural arrangement of this interface does not
depend on vanadia loading and is persistent in samples
even after extraction of excess V2O5 and soluble components [101]. It also remains mostly intact during catalytic
reactions and under water adsorption.
The structure of vanadium sites in this strongly bound
vanadium phase was rst studied with 51V NMR by Shubin
et al. [100]. They have found that these vanadium species
are characterized by unusually large 51V quadrupolar coupling constants, in the range from 14 to 16 MHz, whereas
the principal components of the CSA tensor are similar
to those found in bulk V2O5. Examples of the 51V NMR
spectra recorded for the SBV phase formed in VOx/TiO2
catalysts are shown in Fig. 44.
It is interesting, that earlier theoretical studies also predicted a possibility of strong bonding and intergrowth
between crystalline phases of anatase and vanadium pentoxide [256]. According to these calculations and 51V
NMR data, vanadium sites in the SBV phases are in distorted octahedral oxygen coordination similar to V2O5;
yet the very large quadrupolar coupling constant, almost

179

an order of magnitude of that found in V2O5, indicates


considerable structural distortions.
Strongly bound vanadium has later been found in samples prepared not only by spray-drying, but also in samples
prepared by many other techniques, including grafting,
impregnation and mechano-chemical activation. However,
51
V NMR parameters found for SBV in all these systems
may vary, most likely due to impurities in anatase, specics
of the preparation procedure, or in some cases due to the
presence of water. The dierences are found mostly in the
quadrupolar coupling constant [257], indicating that dierences among samples exist in the long-range order, while
the short-range order reected in the CSA parameters is
usually close and is similar to V2O5. The most symmetric
SBV species are found in samples prepared by grafting.
In VOx/TiO2 samples prepared via ball milling followed
by thermal treatment the parameters of the stacking geometry are quite dierent from all other studied systems. The
SBV phases in these samples also lack the coherent tting
between lattices of the dierent phases with rising of the
strain dislocations at the boundary. 51V NMR experiments
of these samples revealed two dierent types of octahedrally coordinated vanadium species [258]. More distorted
species are formed predominantly during milling, whereas
more symmetric species are found after thermal treatment.
In both cases the distortion is less axial than in the bulk
V2O5.
The structure of SBV species also depends on the support used. Thus on tetragonal ZrO2 a variety of sites can
be found but all of them have a tetrahedral coordination.
51
V 3QMAS NMR spectra indicate the presence of at least
10 non-equivalent vanadium sites in these samples
(Fig. 45).
Strongly bound vanadium species are also formed on
monoclinic ZrO2, and Nb2O5, but not on SiO2.

after
reaction
initial

-1000

-500 -600 -700

Id, ppm

-800

-600
-400
-200

4000

2000

0
-2000
, ppm

-4000

-6000

51

Fig. 44. V MAS NMR spectra recorded at 9.4 T for strongly bound
vanadium sites in VOx/TiO2 catalysts prepared by spray-drying. (1)
Experimental spectrum with the full set of spinning sidebands. (2)
Simulated spectrum with the following NMR parameters: CQ = 4.1 MHz,
gQ = 0.12, dr = 520 ppm, gr = 0.1, diso = 634 ppm, dispersion of distribution DgQ = 2 kHz. Inset shows isotropic lines in the spectra of the initial
catalyst and the catalyst tested in the DeNOx catalytic reaction.

ppm
-500

-550

, ppm

-600

-650

Fig. 45. 51V 3QMAS NMR spectrum of a VOx/ZrO2 catalyst prepared by


impregnation. The spectrum was obtained at 9.4 T.

180

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

x2.3

3
4

5
6
1000

-1000

-2000

, ppm
Fig. 46. 51V MAS NMR spectra recorded at 9.4 T for VOx/Al2O3
catalysts prepared by impregnation with vanadyl oxalate. (1) Spectrum of
the initial 25 wt% V2O5/Al2O3 sample. (2) Spectrum of the sample after
extraction of soluble components. Also shown sub-spectra representing (3)
V2O5, (4) V10 O28 6 , (5) AlVO4, and (6) SBV species.

Large vanadium-containing species like decavanadates,


V10 O28 6 , and aggregates of V2O5, are formed on the surface either when vanadium loading exceeds the monolayer
coverage, or when dispersion of vanadia is not suciently
high. Decavanadates can only be unambiguously detected
in 51V MAS NMR spectra recorded under very fast
magic-angle spinning, in excess of 3035 kHz. At ambient
conditions V10 O28 6 species were detected in almost all
studied VOx/Al2O3 catalysts with vanadium content
exceeding the monolayer coverage. One such example of

a 51V MAS NMR spectrum recorded for a VOx/Al2O3 catalyst is shown in Fig. 46. Several vanadium sites can be
identied in this spectrum by deconvoluting it into subspectra corresponding to surface tetrahedral sites,
V10 O28 6 species, V2O5 and AlVO4. It is noteworthy, that
after extraction of soluble components, the AlVO4 phase
is still retained in the sample together with SBV sites
(Fig. 46, spectrum 2).
When V2O5 is formed on the surface, its structure is
always distorted due to structural defects in the surface.
This can be seen particularly well by observing the broadening of the satellite transitions as shown in the static spectra in Fig. 47. The parameters of the magnetic shielding are
practically not aected.
Binary phases may form under certain conditions upon
interaction of vanadium with the substrate. For example,
when Al2O3, Nb2O5 or ZrO2 are used as supports, the formation of AlVO4, NbVO5, and ZrV2O7 can be observed
(Figs. 46 and 48) [259]. Titania and silica supports do not
normally form binary compounds with vanadium.
Third component, i.e. an impurity in the starting materials or when a modier agent is added, can considerably
alter the structure of vanadium species. For example,
when phosphorus is present during spray-drying, the
structure of the resulting SBV species is aected. Here
phosphorus is often being incorporated into the nal
product with the molar ratio of V/P/Ti close to 1:1:1.
Vanadium in these VPTi species is found in the tetrahedral coordination typical for isolated or weakly associated tetrahedra (Fig. 49) [180]. New compounds have
also been found in VOx/TiO2 catalysis modied with
sodium [260].
Bulk vanadium-containing catalysts based on bismuth
vanadate, Bi4V2O11, are promising candidates for applications in the oxidative coupling of methane. Bismuth vanadate itself shows unusually high anionic conductivity

1
1
2

2
3
4

3
-1000

-1500

-1000

-1500

1
1
2
3
4

2
3

-1000
, ppm

-2000

-1000
, ppm

-2000

Fig. 47. (A) Static 51V NMR spectra recorded at 9.4 T for (1) (V2O5WO3)/TiO2 catalysts after catalytic reaction, (2) starting (V2O5WO3)/TiO2 catalyst,
and (3) V2O5. Inset at the top shows ne details in the satellite transitions. (B) Simulated static 51V NMR spectra with the distribution of the quadrupolar
p
constant DCQ at around 799 kHz according to the Gaussian function gx 1=D 2p expx2 =2D2Q . (1) DQ = 100 kHz, (2) DQ = 75 kHz, (3)
DQ = 50 kHz, (4) DQ = 0 kHz. Inset at the top shows ne details in the satellite transitions.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

181

Fig. 48. 51V MAS NMR spectra recorded at 9.4 T for (A) VOx/Nb2O5 and (B) VOx/ZrO2 catalysts. Vanadium loading was 2 wt% in (1), 4 wt% in (2) and
8 wt% in (3). The isotropic lines are marked with asterisks. Reproduced with permission from Ref. [259].

-500
-1000
, ppm

-1500

Fig. 49. 51V NMR spectra obtained at 9.4 T for a 5%P2O5/10%V2O5/TiO2


catalyst prepared by spray drying. (1) Stationary sample. (2) 10 kHz MAS.
(3) 30 kHz MAS.

sought for advanced composite materials. Not surprisingly,


these vanadium systems are currently under scrutiny often
involving solid-state 51V NMR.
According to 51V NMR data reported by Delmaire et al.
[189] there are at least three dierent vanadium sites in
a-Bi4V2O11, two of these sites are tetrahedral and one is
in bipyramidal trigonal or has a distorted octahedral oxygen coordination. The majority of the vanadium (about
70%) occupies regular tetrahedral sites and the rest is found
in bipyramidal trigonal or distorted octahedral sites.
Reduction of a-Bi4V2O11 under a ow of hydrogen gas at
330 C results in progressive conversion of V5+ into V4+.
At rst a broad signal due to paramagnetic V4+ ions

appears in the 51V NMR spectra underneath rather narrow


resonances from the starting material a-Bi4V2O11. Following further reduction a new resonance at 1447 ppm with
multiple MAS sidebands becomes apparent. This resonance is dominant in Bi6V3O16, the nal product of bismuth vanadate reduction (Fig. 50). In this compound the
ratio V5+/V4+ = 2/1. The room-temperature crystal structure of Bi6V3O16 can best be described as ribbons of
V3O10 units in which one V4+O6 octahedron shares equatorial corners with two V5+O4 tetrahedra intercalated
between sheets. The paramagnetic shift observed in 51V
NMR spectra of Bi6V3O16 was attributed to strong interactions between V5+ ions and unpaired electrons of V4+ ions.
Summarizing the results outlined above we would like to
once more accentuate the idea that the 51V NMR technique
is a unique research tool indispensable for studying vanadia catalysts and related materials. The possibility to closely follow the formation and transformations of
vanadium-containing centers during synthesis and under
catalytic conditions makes it possible to purposefully
design catalysts with well-dened and, more importantly,
desirable properties.
9.2. Applications of solid-state

93

Nb NMR

While 51V NMR nds most of its applications in catalysis, 93Nb NMR is gaining in popularity in studying materials for advanced electronic devices and semiconductors.
In these challenging systems 93Nb solid-state NMR demonstrates a great potential to provide both qualitative and
quantitative information about chemical environments,
cation ordering, and the motional behavior of cations. Several 93Nb NMR examples in solids have already been discussed above, and some applications have recently been
reviewed [55].

182

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

Fig. 50. 51V MAS NMR spectra recorded at 9.4 T for Bi4V2O11 reduced under a hydrogen gas ow at 330 C for (1) 10 h and (2) 16 h. The isotropic lines
of dierent vanadium sites are marked as 1, 2, 3. Reproduced with permission from Ref. [189].

9.2.1. Characterization of ferroelectrics


Niobia-based ferroelectric materials are currently the
subject of intensive research, with several groups reporting
their 93Nb NMR ndings, usually complemented by results
from other research techniques, including neutron and Xray-diraction experiments.
Thus polycrystalline solid-solution relaxor ferroelectrics
Pb(Mg1/3Nb2/3)O3 (PMN) and (1x)Pb(Mg1/3Nb2/3)O3/
xPbTiO3 (PMN/PT) have been studied using 93Nb MAS
and nutation NMR at 9.4 and 14.1 T [47]. The authors
Table 10
93
Nb resonances found in PMN materials from
NMR experiments [47]
Site

1
93

dobs ( Nb, MAS),


(ppm)
CQ (93Nb, nutation),
(MHz)
B-site symmetry
Assignment

900
(narrow)
<0.8
Cubic
Nb(OMg)6

93

Nb MAS and nutation


3

954 to 980 (broad)


17

>62

Tetragonal
Rhombic
Nb(ONb)6x(OMg)x
(x = 15)

obtained accurate atomic-level information about the local


structure and chemical disorder of Nb cations occupying
B-sites. In these and similar systems 93Nb MAS and nutation NMR spectra usually show three 93Nb resonances
with drastically dierent quadrupolar coupling constants,
assigned to Nb sites with cubic, axial and rhombic local
symmetry of the neighboring Mg/Nb congurations. Typical 93Nb NMR parameters for these three sites are summarized in Table 10. Three 93Nb resonances in PMN were also
observed by Prasad et al. [48], and by Cruz et al. [49].
According to Fitzgerald et al. [47] the Nb sites corresponding to sharp 93Nb signals at 902 ppm are most
likely associated with cubic Nb(OMg)6 congurations in
Mg-rich ordered-domain regions. Two broader 93Nb
NMR resonances with intermediate and high CQ values
were assigned to two types of lower symmetry NbO6 octahedra with local congurations of Nb(ONb)6x(OMg)x
type, where x may vary from 1 to 5. One set of congurations has axial local symmetry, and another has lower
rhombic symmetry.
93
Nb NMR allows the monitoring of compositional
changes in PMN/PT solid-solutions. At low PT content,

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

Ti ions substitute both Mg and Nb in the B-sites of PMN.


At increased PT content (>10 mol%) the simultaneous
presence of three cations Mg, Ti, and Nb randomly
occupying B-sites leads to increased structural disorder of
the B-sites. This causes the sharp 93Nb signal from
near-cubic Nb(OMg)6 sites to almost disappear, and
further changes to occur in the lineshapes from the
other two resonances assigned to axial and rhombic
Nb(ONb)6x(OMg)x sites in Nb-rich regions of PMN. Ti
substitution for Nb in both Mg-rich and Nb-rich regions
of PMN was found to be nonselective. At the same time
there are distinct dierences between Nb B-sites located
in Mg- and Nb-rich regions in PMN and PMN/PT materials at low PT contents [47].
A comparative study of lead-based perovskite relaxor
ferroelectrics, Pb(Mg1/3Nb2/3)O3 (PMN), substituted by
Zr, Sc, and Ba (PZN, PSN, PBN correspondingly), have
recently been reported to show dierences in the local
chemical environments, in the relative degree of B-site
ordering, and in the motional behavior of the B-site cations [51]. 93Nb MAS NMR spectra provided conclusive
evidences for signicant B-site disorder in PMN, while
B-site ordering was found in scandium-substituted
(PSN) materials. In agreement with earlier studies 93Nb
MAS NMR spectra of PMN indicate the presence of several resonances with very dierent CQ values, which were
assigned to dierent types of Nb B-sites as in pseudocubic, axial and rhombic symmetry. At the same time in
PSN there was found to be a predominantly single Nb site
with an intermediate value of CQ, which was assigned to
axially distorted B-sites. Substitution by Ba results in
increased B-site ordering in the Mg-rich and Nb-rich nanodomains of PMN.
Scandium substituted ferroelectrics (PSN) have also
been studied by Blinc et al. [52,261]. 93Nb NMR spectra
presented there suggest a homogeneous structure of the
PMN and PSN phases studied. Both structures can be considered within a spherical random bond-random eld
model [262] where disorder of the Pb2+ sites is represented
by the order parameter described as a continuous vector of
variable length. While studying similar PSN materials,
Laguta et al. [263] reported excellent agreement between
data obtained from neutron and X-ray-diraction experiments and via multinuclear solid-state NMR. In this work
207
Pb, 45Sc, and 93Nb NMR spectra of partially ordered
PSN were obtained at temperatures between 77 and
420 K to reveal a rst-order phase transition at 360 K. This
transition was particularly straightforward to detect in
somewhat narrower 45Sc and 207Pb NMR spectra originating from chemically ordered regions of the crystal, but not
in the broader 93Nb NMR spectra.
9.2.2. Silicates
Silicates containing niobium are quite often studied with
93
Nb NMR, usually in combination with other nuclei
NMR, for example, 29Si. Thus the structure of nenadkevichite minerals, Na0.9K0.06Ca0.03 Nb0.7Ti0.3Si2O6.7(OH)0.3

183

2H2O, was characterized by multinuclear 23Na, 29Si, and


93
Nb high-resolution solid-state NMR [41]. In agreement
with the crystal structure reported for this mineral, 93Nb
solid-state NMR suggests that both titanium and niobium
are in a distorted octahedral coordination. Due to stronger
dipolar interactions, replacing low-abundant 47Ti and 49Ti
with 93Nb having high natural abundance results in a signicant decrease in the 29Si spinspin (T2) NMR relaxation
times.
Static 93Nb NMR spectra have been reported for a series
of (Nb2O5)x(SiO2)1x materials prepared via a sol-gel technique [42]. The authors suggested that niobium is incorporated into silica as a four-coordinated species. This
conclusion is somewhat questionable, however, since not
only are the reported 93Nb NMR spectra very similar to
those normally found for Nb2O5, but also peaks corresponding to bridging NbONb groups can be clearly seen
in the 17O NMR spectra. The reported isotropic 93Nb
NMR chemical shifts values of about 1200 ppm are also
close to those characteristic of six-coordinated niobium. It
appears that niobium was indeed incorporated into SiO2,
but not as the isolated tetrahedral Nb species the authors
stipulated.
A combination of 93Nb and 29Si solid-state NMR
and X-ray diraction data has provided a correlation
between NMR parameters and the local structure in
Rb4(NbO)2(Si8O21) silicate [126]. For the rst time heteronuclear two-bond J-coupling between a quadrupolar
nucleus (93Nb) and a spin-1/2 nucleus (29Si) was observed
in the solid state. This work has opened up new opportunities for investigating possible relationships between
J-coupling and structural parameters in solids.
9.2.3. Miscellaneous applications
Among some other recent applications of 93Nb NMR in
the solid state we would like to bring our readers attention
to research by Du et al. [54] who applied modern NMR
techniques to study local structures and oxygen/uorine
ordering in Cdpy4NbOF5 (py = C5H5N) and [pyH]2[Cdpy4(NbOF5)2] compounds. Their 93Nb NMR spectra were
acquired at ultrahigh magnetic eld of 19.6 T and MAS
under ultrahigh spinning speeds (43 kHz). The 93Nb
NMR spectra of both compounds are dominated not only
by quadrupolar but also by the chemical shift interactions,
consistent with a highly asymmetric Nb environment
(Fig. 51). To extract the nuclear quadrupolar coupling constant CQ, the asymmetry parameter gQ, and the isotropic
chemical shift diso in the presence of the large CSA, the
authors applied the original methodology as described in
detail in Section 3.5. The values of the 93Nb NMR parameters found for these two compounds are summarized in
Table 7.
More recently Lo et al. [264] reported 93Nb NMR data
for a series of half-sandwich niobium metallocenes
Cp 0 Nb(I)(CO)4 and CpNb(V)Cl4. They have found the
93
Nb chemical shifts for Nb(I) metallocenes ranging from
1900 to 2050 ppm, and for Nb(V) metallocenes from

184

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

Fig. 51. (A) 93Nb NMR spectra of Cdpy4NbOF5. (B) 93Nb NMR spectra of [pyH]2[Cdpy4(NbOF5)2]. (1) Stationary samples at 8.5 T. (2) 20 kHz MAS at
8.5 T. (3) 43 kHz MAS at 19.6 T. Reproduced with permission from Ref. [54].

600 to 715 ppm. The observed dierence in the chemical


shift values for two Nb oxidation states was explained by
shielding eects resulting from the Cp 0 -metal p-bonding
interactions. Although metallocenes are not the subject of
this review, this paper is a good example of a well-executed
and comprehensive solid-state 93Nb NMR study.
Another interesting recent case involving paramagnetic
vanadium species was reported by Ponce et al. [43]. While
studying 51V and 93Nb NMR spectra of HxNbVO5
(x = 1.4) bronze, they found no dierences in the chemical
shift and the line width of the central transition in this compound with respect to those observed in NbVO5. From
thermodynamic calculations the charge transfer from
hydrogen to the host lattice in HxNbVO5 appears to be
complete. Thus the formal oxidation state of vanadium in
this bronze should be V3.5+ and the abundance of V4+
and V3+ paramagnetic centers should have prevented
observation of 51V and 93Nb resonances. The reported
NMR results are dicult to explain within the thermodynamic model of the complete charge transfer, unless some
electron-pairing occurs. Solving this problem may prove
to be a challenging task requiring additional experimental
and computational techniques to be involved.
93
Nb NMR spectra of a mixed-cation KTa(1x)NbxO3
(KTN) single crystal have been recorded by Runkel et al.
[265]. The spectra consist of two components, a sharp central +1/2 M 1/2 transition along with an unresolved
background from quadrupolar-induced rst-order satellites. These spectra were interpreted as being from Nb ions
occupying o-center positions rather than from those in the
high symmetry central perovskite site. The angular dependence of the second moments of the satellite background
further showed that these distortions are of rhombohedral
symmetry, i.e. Nb ions are indeed displaced along the [111]
body diagonals. Another recent single-crystal 93Nb NMR
study involved measurements of the 93Nb EFG parameters
in Nb-doped single TiO2 crystals [266].
93
Nb and 27Al NMR studies of composites between partially hydrolyzed aluminum cations and layered calcium

niobate perovskites suggest the presence of structurally


unaltered Ca2Nb3O10 layers in these materials with interlayer regions occupied by polymeric Al hydroxylcations
[36].
9.3. Multinuclear solid-state NMR in vanadia and niobia
catalysts supported on Al2O3
This example of solid-state 51V and 93Nb NMR is perhaps one of the best to fully illustrate the advantages of this
technique for studying complex multi-component systems
having tremendous practical importance. The multinuclear
NMR approach, involving simultaneous use of both 51V
and 93Nb nuclei, has allowed us to characterize these catalysts in great detail, and to provide a basis for improving
their performance in real practice. In the next few paragraphs we will summarize results of several years of
research and the eorts of many of our colleagues.
9.3.1. Vanadia sites in VOx/Al2O3
As it has been mentioned above, several types of vanadia sites can be identied by 51V NMR in supported vanadia catalysts. In VOx/Al2O3 systems these include weakly
and strongly bound surface species, AlVO4, large V10 O28 6
anions, and distorted V2O5 [267]. The relative amounts of
these sites on the surface depend on the preparation procedure, the total vanadium concentration, humidity, and pH
of the surface. What is important in this context is that
each of these sites has its own distinct 51V NMR signature
(Fig. 52).
Tetrahedral VO4 species weakly bonded with the surface
via one or two bonds have similar static and MAS 51V
NMR spectra with isotropic shifts found in the range from
540 to 580 ppm. The NMR spectra of stronger bound
VO4 species are shifted upeld and usually consist of
broader isotropic lines at 620 ppm. Large vanadia polyanions V10 O28 6 on alumina surface are characterized by
static 51V NMR spectra typical for an axially symmetric
chemical shielding anisotropy. At high MAS spinning

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

185

VOx /Al2O3
V2O5, 10.8 wt.%

V2O5

AlVO4

V10O286-

O
V
OOO

400

-500

-1000

-1500

, ppm
0

-500

-1000

-1500

51

Fig. 52. (Left) V MAS and static NMR spectra at 9.4 T of dierent vanadium species formed in VOx/Al2O3 catalysts as indicated. (Right) Relative
content of dierent vanadium species in a sample containing 10.8 wt% of V2O5. The experimental 51V MAS NMR spectrum of this sample is shown on the
top.

speeds of 35 kHz these spectra can be resolved into three


resonances with isotropic shifts diso of 420, 490, and
510 ppm. Static and MAS 51V NMR spectra of the
AlVO4 phase consists of three relatively narrow overlapping lines from three non-equivalent vanadium sites in tetrahedral oxygen coordination with an isotropic shifts diso
of 668, 747, and 780 ppm. The V2O5 phase on the surface is identied by an axially symmetric line in the static
51
V NMR spectra with d^ = 310 ppm, di = 1250 ppm.
Under MAS conditions this spectrum transforms into a
single set of narrow spinning sidebands with isotropic shift
of 610 ppm. Some distribution of NMR parameters
reects a defect structure of this phase on the surface.
At low vanadium loading, only tetrahedral species
(weakly and strongly bound to the surface) are formed
on the alumina surface. With increasing vanadium content
the relative concentration of these sites decreases, at the
same time decavanadate V10 O28 6 species are becoming
prevalent. Further increase in the vanadium content results
in formation of AlVO4 and V2O5 phases. The relative content of each type of vanadia species can be determined by
analyzing the corresponding 51V NMR spectra as shown
in Fig. 52 for a sample containing 10.8 wt% of V2O5.
Upon dehydration the surface vanadia species in
VOx/Al2O3 undergo certain transformations, which can
easily be monitored with 51V NMR. Thus tetrahedral VO4
species (weakly and strongly bound to the surface) become
more distorted in the dehydrated state. Decavanadate species are converting into signicantly distorted tetrahedral
sites of both types, with the ratio between them depending
on the initial content of V10 O28 6 , i.e. the share of strongly
bound species increases with increase in V10 O28 6 . Both
AlVO4 and V2O5 phases adopt a more regular structure
upon dehydration. The relative ratio of dierent vanadium

sites in the dehydrated samples depends on the relative ratio


of V sites in the corresponding hydrated samples.
According to 27Al MAS NMR recorded for the same
catalysts there are no strong interactions between V and
Al in VOx/Al2O3 samples at dierent vanadium loadings.
Only at the very high vanadium content, when the AlVO4
phase is formed, the new 27Al MAS NMR line appears
at 10 ppm corresponding to AlVO4 species. The latter
can be better seen in corresponding 3Q MAS 27Al NMR
spectra.
Interaction of vanadia with alumina in VOx/Al2O3 can
also be monitored by 1H MAS NMR of hydroxyl groups.
Vanadia species interact with the surface of Al2O3 via
bridged OH groups (d1H 1.6 ppm) and terminal OH groups
(d1H 0.2 ppm). At the same time hydrogen-bonded
hydroxyl groups (d1H 7.4 ppm) and bulk OH groups
(d1H 3.7 ppm) remain largely intact. These data indicate
high dispersion of vanadia species over the Al2O3 surface.
At the same time there is no noticeable change in the bulk
Al2O3 structure.
9.3.2. Niobia sites in NbOx/Al2O3
As discussed above and by Lapina et al. [55] the 93Nb
NMR chemical shift is sensitive to Nb coordination in
oxide materials. Representative static 93Nb NMR spectra
for a series of NbOx/Al2O3 samples recorded at 21.1 T
are shown in Fig. 53. This is perhaps the rst time such
spectra have been reported for NbOx/Al2O3 catalysts. All
recorded spectra are rather broad with no narrowing
observed under MAS. The broadening is most likely
caused by both the quadrupolar interaction and by the
chemical shift distribution. When recorded at dierent
magnetic elds, the line width measured in parts per million
scales inversely with the magnetic eld strength, but only

186

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

500

-1000

-3000

, ppm
Fig. 53. 93Nb NMR spectra of stationary NbOx/Al2O3 samples recorded
at 21.1 T. (1) 4% Nb2O5/Al2O3. (2) 8% Nb2O5/Al2O3. (3) 16% Nb2O5/
Al2O3.

linearly and not as a square function expected for pure


quadrupolar broadening (see above). At low niobia loading
the coordination number of Nb sites in NbOx/Al2O3 is
close to six, i.e. as found in octahedral NbO6 species. At
higher Nb loadings a new 93Nb NMR resonance can be
seen in the spectra recorded at 21.1 T. By the value of its
chemical shift this resonance can be attributed to sevenor eight-coordinated Nb sites. At lower magnetic elds

V2O5

O
V
OOO

the spectra are too broad for this new resonance to be


resolved.
As was mentioned above, in 27Al MAS NMR spectra of
VOx/Al2O3 catalysts we normally do not see strong interactions between vanadia and alumina. However, this is not
the case for niobia catalysts. In similar 27Al MAS NMR
spectra recorded for NbOx/Al2O3 samples there was
detected a very characteristic 27Al resonance attributed to
ve-coordinated Al sites [268]. The intensity of this line
increases with Nb content. It is reasonable to suggest stronger interactions of niobia with alumina, with possible
incorporation of Nb into alumina. The Nb coordination
number increases with loading, thus indicating interactions
among neighboring Nb species, and this supports the
island model of niobia deposition on the alumina surface.
Additional support for the island model can be found in
1
H MAS NMR spectra recorded for the same NbOx/Al2O3
samples. Supported niobia species indeed interact with the
surface hydroxyl groups as vanadia species do. However, at
similar surface concentrations of niobia and vanadia the
content of residual OH groups is always considerably
higher in NbOx/Al2O3 than in VOx/Al2O3 samples, apparently due to higher dispersion of vanadia species on the surface, while niobia species tend to form aggregates.
9.3.3. Niobiavanadia species in (NbV)Ox/Al2O3
Vanadia and niobia can simultaneously be deposited
onto alumina to prepare mixed (NbV)Ox/Al2O3 samples.
The nature of niobia and vanadia species in such systems is

Nb2O5
NbAl

AlVO4
Al2O3

NbAl

Al2O3
High niobia coverage
O
V
OOO
V or Nb loading

V2O5

Al2O3

O
V
OOO

O
V
OOO
NbAl

NbAl

NbAl

O
V
OOO

O
V
OOO
NbAl

Al2O3

Al2O3

Low niobia loading

Low niobia loading

10

Al2O3
O
V
OOO

O
V
OOO

O
V
OOO
Vanadia feels niobia, but niobia is not affected by vanadia

Al2O3

Fig. 54. Surface species formed in VOx/Al2O3, NbOx/Al2O3, and (NbV)Ox/Al2O3 catalysts.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

of great interest, since they often show a synergetic catalytic behavior by exceeding the performance of individual
single-component VOx/Al2O3 and NbOx/Al2O3 catalysts.
The following conclusions about (NbV)Ox/Al2O3 can be
drawn from multinuclear NMR experiments.
First of all, the new 51V NMR line with the isotropic
chemical shift of 650 ppm appears in 51V NMR spectra
in the presence of niobia. The relative intensity of this line
increases with niobia loading, and becomes even more pronounced in dehydrated samples. It is reasonable to attribute this line to (NbV)Ox species, though it is not yet
NbVO5. At the same time there are only very minor
changes in the corresponding static 93Nb NMR spectra.
27
Al MAS NMR spectra recorded for (NbV)Ox/Al2O3
are similar to those for NbOx/Al2O3, with no strong interactions detected between vanadia species and the alumina
support even at high Nb and V loadings. An almost linear
decrease in the surface concentration of OH groups with
(NbV)Ox loading is observed in the 1H MAS NMR
spectra.
All these NMR ndings can be explained by the island
model introduced above for the NbOx/Al2O3 system. First,
the niobia precursors interact strongly with the alumina
surface forming islands of mixed (NbAl) oxides. The vanadia precursors then interact with both oxides and disperse
randomly on the surfaces of Al2O3 and the mixed (NbAl)
oxide islands. Schematically this is shown in Fig. 54.
Thus, the multinuclear NMR approach has allowed us
in this case not only to identify various surface species in
these catalysts, but also to suggest the most likely model
of the surface structure. Testing these catalysts in catalytic
reactions would provide further input in understanding
structureproperties relationships in these complex
systems.
10. Conclusions
We have reviewed practical aspects of solid-state NMR
spectroscopy of Group VB elements pertaining to studying
a large variety of oxide materials. Some of these materials
already have tremendous importance in large-scale industrial applications, as for example, vanadia-based systems
in heterogeneous catalysis, other materials are currently
being investigated for possible future uses in electronics
and advanced composites. Solid-state NMR oers a unique
and indispensable tool for learning in great detail about the
structure and the underlying properties of these materials.
While 51V and 93Nb NMR spectroscopy has already established itself as a popular research vehicle, there is no doubt
that 181Ta NMR is also about to show its great potential.
The availability of advanced NMR hardware, including
ultrahigh magnetic elds and ultrahigh-speed MAS probes
will make 181Ta NMR experiments on solids feasible and
will further expand the borders of 51V and 93Nb NMR
applications. We have compiled perhaps the most comprehensive database to date on 51V and 93Nb NMR parameters in oxide materials, yet we understand that this

187

database is destined to grow as more and more data


become available every day. We hope this work will contribute to ever continuing progress in development of novel
solid-state NMR techniques for quadrupolar nuclei. Modern NMR spectroscopy combined with other complementary spectroscopic and computational approaches not
only advances our fundamental knowledge in material science and related disciplines, but also helps us to develop
better technologies and to discover new ways of improving
everyday life.
Acknowledgements
Our research in this eld has been generously supported
by the Russian Foundation for Basic Research, most recently by a RFBR Grant 07-03-00695-a. This review would
not be possible without our many collaborators, coauthors,
and colleagues to whom we are extremely grateful. We dedicate this review to the memory of our dear teacher and
friend the late Prof. V.M. Mastikhin. We thank Prof. P.
Bodart (Universite des Sciences et Technologies de Lille),
Dr. K.V. Romanenko (Boreskov Institute of Catalysis,
Novosibirsk), Prof. J.-B. Espinose (ESPCI, Paris), Dr.
Z.H. Gan (NHMF Laboratory, Florida), Prof. J.-P.
Amoureux (Universite des Sciences et Technologies de
Lille), and Prof. M. Guelton (Universite des Sciences et
Technologies de Lille) among others for their expertise in
carrying out some of the experiments and for helpful discussions. Many individual vanadium and niobium compounds have been skillfully synthesized by Prof. V.N.
Krasilnikov (Institute of Solid State Chemistry, Ekateribnurg), Prof. M.G. Zuev (Institute of Solid State Chemistry,
Ekaterinburg), Prof. V.L. Volkov (Institute of Solid State
Chemistry, Ekaterinburg), and Prof. V.E. Fedorov (Nikolaev Institute of Inorganic Chemistry, Novosibirsk). Research on catalytic systems have been done over many
years in close collaboration with Prof. M. Bonares (Instituto de Catalisis y Petroleoquimica, Madrid), Prof. B. Grzybowska (Institute of Catalysis and Surface Chemistry,
Krakow), Prof. I. Wachs (Lehigh University, Bethlehem,
USA), Prof. Dr. H. Knozinger (Ludwig-Maximilians-Universitat Munchen), Prof. R. Fehrmann (Technical University of Denmark, Lyngby), Prof. Z. Sobalik (Heyrovsky
Institute of Physical Chemistry, Prague), Dr. V.M. Bondareva (Boreskov Institute of Catalysis, Novosibirsk), Dr.
L.G. Pinaeva (Boreskov Institute of Catalysis, Novosibirsk), Prof. G.A. Zenkovets (Boreskov Institute of Catalysis, Novosibirsk), and Dr. L.G. Simonova (Boreskov
Institute of Catalysis, Novosibirsk). Some NMR spectra
at 21.1 T were recorded at the Canadian National Ultrahigh Field NMR Facility for Solids (Ottawa, Canada), a
national research facility funded by the Canada Foundation for Innovation, the Natural Sciences and Engineering
Research Council of Canada, the Ontario Innovation
Trust, Recherche Quebec, the National Research Council
Canada, and Bruker BioSpin, and managed by the University of Ottawa (www.nmr900.ca).

188

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

References
[1] G.K. Boreskov, Catalysis in the Production of Sulfuric Acid,
Goskhimizdat, Moscow, 1954 (in Russian).
[2] G.K. Boreskov, Heterogeneous Catalysis, Nauka, Moscow, 1987 (in
Russian).
[3] D.J. Hucknall, Selective Oxidation of Hydrocarbons, Academic
Press, London, 1974.
[4] B.E. Leach, Applied Industrial Catalysts, Academic Press, NewYork, 1983.
[5] A. Andersson, J. Catal. 69 (1981) 465.
[6] G. Centi, Appl. Catal. A 147 (1996) 267.
[7] T. Machej, P. Ruiz, B. Delmon, J. Chem. Soc., Faraday Trans. 1
(1990) 731.
[8] F. Triro, I. Pasquon, J. Catal. 12 (1968) 412.
[9] O.V. Krylov, Catalysis by Nonmetals, Academic Press, New York,
1970.
[10] K. Kenneth, C.-P. Joao, C.C. Debbie, Vanadium: The Versatile
Metal, American Chemical Society, Washington, 2007.
[11] M.A. Banares, I.E. Wachs, R.M. Martin-Aranda (Eds.), in: Fourth
International Symposium on Group Five Compounds, Toledo,
Spain, 912 April 2002, UNED, Madrid, 2002.
[12] I.E. Wachs (Ed.), in: Fifth International Symposium on Group Five
Compounds, Hancock, Massachusetts, USA, 1821 May 2005,
Lehigh University, USA, 2005.
[13] K. Tanabe, Catal. Today 78 (2003) 65.
[14] W.D. Knight, V.W. Cohen, Phys. Rev. 76 (1949) 1421.
[15] R.E. Sheri, D. Williams, Phys. Rev. 82 (1951) 651.
[16] L.H. Bennett, J.I. Budnick, Bull. Am. Phys. Soc. 4 (1959) 417.
[17] J.L. Ragle, J. Chem. Phys. 35 (1961) 753.
[18] S.D. Gornostansky, G.V. Stager, J. Chem. Phys. 46 (1967) 4959.
[19] S.D. Gornostansky, C.V. Stager, J. Chem. Phys. 48 (1968) 1416.
[20] G.E. Peterson, J.R.J. Carruthers, J. Solid State Chem. 1 (1969) 98.
[21] R. Kind, H. Granicher, Solid State Commun. 6 (1968) 439.
[22] R.N. Pletnev, N.A. Gubanov, A.A. Foiev, NMR in Oxide Vanadium Compounds, Nauka, Moscow, 1979 (in Russian).
[23] J.F. Baugher, P.C. Taylor, T. Oja, P. Bray, J. Chem. Phys. 50 (1969)
4914.
[24] S.L. Segel, R.B. Creel, Can. J. Phys. 48 (1970) 2673.
[25] R.N. Pletnev, V.A. Gubanov, A.K. Chirkov, Zh. Struk. Khim. 17
(1976) 938 (in Russian).
[26] R.N. Pletnev, Zh. Neorg. Khim. 51 (1977) 2359 (in Russian).
[27] G.E. Peterson, A. Carnevale, J. Chem. Phys. 56 (1972) 4848.
[28] E.N. Ivanova, A.V. Yatseako, N.A. Sergeev, Solid State Nucl.
Magn. Reson. 4 (1995) 381.
[29] J. Blumel, E. Born, T. Metzger, J. Phys. Chem. Solids 55 (1994) 589.
[30] F. Wolf, D. Kline, H.S. Story, J. Chem. Phys. 53 (1970) 3538.
[31] A.F. McDowell, M.S. Conradi, J. Haase, J. Magn. Reson. 119
(1996) 211.
[32] V.M. Mastikhin, O.B. Lapina, V.N. Krasilnikov, A.A. Ivankin,
React. Kinet. Catal. Lett. 24 (1984) 119.
[33] V.M. Mastikhin, O.B. Lapina, L.G. Simonova, React. Kinet. Catal.
Lett. 24 (1984) 127.
[34] O.B. Lapina, V.M. Mastikhin, A.A. Shubin, V.N. Krasilnikov, K.I.
Zamaraev, Prog. Nucl. Magn. Reson. Spectrosc. 24 (1992) 457.
[35] J. Davis, D. Tinet, J.J. Fripiat, J.M. Amarillo, B. Casal, E. RuizHitzky, J. Mater. Res. 6 (1991) 393.
[36] S. Hardin, D. Hay, M. Millikan, J.V. Sanders, T.W. Turney, Chem.
Mater. 3 (1991) 977.
[37] K. Kato, C. Zheng, J.M. Finder, S.K. Dey, Y. Torii, J. Am. Ceram.
Soc. 81 (1998) 1869.
[38] A.H. Munhoz, S. Rodrigues, T. Pinnavaia, Adv. Sci. Technol. 16
(1999) 521.
[39] L.P. Cruz, J.-M. Savariault, J. Rocha, J.-C. Jumas, J.D. Pedrosa de
Jesus, J. Solid State Chem. 156 (2001) 349.
[40] H. Yoshida, H. Nishihara, S. Yokota, M. Ohyanagi, T. Nakaoki, Z.
Naturforsch. A: Phys. Sci. 53 (1998) 309.

[41] J. Rocha, P. Brandao, Z. Lin, A.P. Esculcas, A. Ferreira, M.W.


Anderson, J. Phys. Chem. 100 (1996) 14978.
[42] K.O. Drake, D. Carta, L.J. Skipper, F.E. Sowrey, R.J. Newport,
M.E. Smith, Solid State Nucl. Magn. Reson. 27 (2005) 28.
[43] A.W. Glynn, A.L. Ponce, X. Lin, J.J. Fripiat, Solid State Ionics 84
(1996) 213.
[44] V.M. Mastikhin, O.B. Lapina, Vanadium Catalysts: Solid State
NMR, in: D.M. Grant, R.K. Harris (Eds.), Encyclopedia of Nuclear
Magnetic Resonance, vol. 8, Wiley, New York, 1996, pp. 48924904.
[45] O.B. Lapina, A.A. Shubin, D.F. Khabibulin, V.V. Terskikh, P.R.
Bodart, J.-P. Amoureux, Catal. Today 78 (2003) 91.
[46] A.A. Shubin, O.B. Lapina, D. Courcot, Catal. Today 56 (2000) 379.
[47] J.J. Fitzgerald, S. Prasad, J. Huang, J.S. Shore, J. Am. Chem. Soc.
122 (2000) 2556.
[48] S. Prasad, P. Zhao, J. Huang, J.J. Fitzgerald, J.S. Shore, Solid State
Nucl. Magn. Reson. 14 (1999) 231.
[49] L.P. Cruz, J. Rocha, J.D.P. de Jesus, J.M. Savariault, J. Galy, Solid
State Nucl. Magn. Reson. 15 (1999) 153.
[50] S. Prasad, P. Zhao, J. Huang, J.J. Fitzgerald, J.S. Shore, Solid State
Nucl. Magn. Reson. 19 (2001) 45.
[51] J.J. Fitzgerald, J. Huang, H. Lock, Solid-state 25Mg, 45Sc and 93Nb
MAS NMR studies: local B-site chemical environments and ordering in Pb(Mg1/3Nb2/3)O3 (PMN) and related relaxor ferroelectrics,
in: C. Galassi, M. Dinescu, K. Uchino, M. Sayer (Eds.), Piezoelectric
Materials: Advances in Science, Technology and Applications. vol.
76, NATO Sci. Ser. Kluwer Academic Publishers, Norwell, 2000. pp.
203218.
[52] R. Blinc, A. Gregorovic, B. Zalar, R. Pirc, V.V. Laguta, M.D.
Glinchuk, J. Appl. Phys. 89 (2001) 1349.
[53] L.-S. Du, R.W. Schurko, N. Kim, C.P. Grey, J. Phys. Chem. A 105
(2001) 760.
[54] L.-S. Du, R.W. Schurko, N. Kim, C.P. Grey, J. Phys. Chem. A 106
(2002) 7876.
[55] O.B. Lapina, D.F. Khabibulin, K.V. Romanenko, Z.H. Gan, M.G.
Zuev, V.N. Krasilnikov, V.E. Fedorov, Solid State Nucl. Magn.
Reson. 28 (2005) 204.
[56] D. Rehder, B. Wolf, J. Magn. Reson. 68 (1986) 157.
[57] J.J. Van der Klink, F. Borsa, Phys. Rev. B: Condens. Matter 30
(1984) 52.
[58] K.D. Becker, U. Berlage, J. Magn. Reson. 54 (1983) 212.
[59] J.I. Budnick, H. Lawrence, L.H. Bennett, J. Phys. Chem. Solids 16
(1960) 37.
[60] H. Nishihara, G.A. Scholz, M. Naito, R.F. Frindt, S. Tanaka, J.
Magn. Magn. Mater. 31-34 (1983) 717.
[61] M. Naito, H. Nishihara, S. Tanaka, J. Chem. Phys. C 16 (1983) 387.
[62] E.B. Doering, J.S. Waugh, J. Chem. Phys. 85 (1986) 1753.
[63] S. Rod, F. Borsa, J.J. Van der Klink, Phys. Rev. B: Condens. Matter
38 (1988) 2267.
[64] J. Mason, Multinuclear NMR, Plenum Press, New York and
London, 1987.
[65] P.R. Bodart, J.-P. Amoureux, Y. Dumazy, R. Lefort, Mol. Phys. 98
(2000) 1545.
[66] S.E. Ashbrook, S. Wimperis, J. Magn. Reson. 156 (2002) 269.
[67] J.P. Amoureux, L. Delevoye, S. Steuernagel, Z.H. Gan, S. Ganapathy, L. Montagne, J. Magn. Reson. 172 (2005) 268.
[68] K. Paulsen, D. Rehder, Z. Naturforsch. A37 (1982) 139.
[69] M.A. Habayeb, O.E. Hileman, Can. J. Chem. 58 (1980) 2115.
[70] A. Allerhand, J. Chem. Phys. 52 (1970) 2162.
[71] E.L. Hahn, Phys. Rev. 80 (1950) 580.
[72] P. Manseld, Phys. Rev. 137 (1965) A961.
[73] P.P. Man, E. Duprey, J. Fraissard, P. Tougne, J.-B. dEspinose,
Solid State Nucl. Magn. Reson. 5 (1995) 181.
[74] P.P. Man, Quadrupole couplings in nuclear magnetic resonance, in:
R.A. Meyers (Ed.), Encyclopedia of Analytical Chemistry, Chichester, New York, 2000, pp. 1222412265.
[75] D. Massiot, I. Farnan, N. Gautier, D. Trumeau, A. Trokiner, J.P.
Coutures, Solid State Nucl. Magn. Reson. 4 (1995) 241.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191
[76] G. Wu, S. Dong, Solid State Nucl. Magn. Reson. 20 (2001) 100.
[77] F.H. Larsen, J. Skibsted, H.J. Jakobsen, N.C. Nielsen, J. Am. Chem.
Soc. 122 (2000) 7080.
[78] A.S. Lipton, J.A. Sears, P.D. Ellis, J. Magn. Reson. 151 (2001) 48.
[79] R. Lefort, J.W. Weinch, M. Pruski, J.-P. Amoureux, J. Chem. Phys.
116 (2002) 2493.
[80] G.M. Bowers, A.S. Lipton, K.T. Mueller, Solid State Nucl. Magn.
Reson. 29 (2006) 95.
[81] F.H. Larsen, H.J. Jakobsen, P.D. Ellis, N.C. Nielsen, J. Magn.
Reson. 131 (1998) 144.
[82] I. Hung, R.W. Schurko, J. Phys. Chem. B 108 (2004) 9060.
[83] K.J. Ooms, K.W. Feindel, V.V. Terskikh, R.E. Wasylishen, Inorg.
Chem. 45 (2006) 8492.
[84] A. Samoson, E. Lippmaa, Phys. Rev. B: Condens. Matter 28 (1983)
6567.
[85] E. Oldeld, R.A. Kinsey, B. Montez, T. Ray, K.A. Smith, J. Chem.
Soc. Chem. Commun. (1982) 254.
[86] A. Samoson, Extended magic-angle spinning, in: D.M. Grant, K.
Harris (Eds.), Encyclopedia of Nuclear Magnetic Resonance 9,
Wiley, Chichester, 2002, pp. 5964.
[87] A. Samoson, T. Tuherm, J. Past, J. Magn. Reson. 149 (2001) 264.
[88] M. Ernst, A. Samoson, B.H. Meier, Chem. Phys. Lett. 348 (2001)
293.
[89] A. Samoson, T. Tuherm, Z.H. Gan, Solid State Nucl. Magn. Reson.
20 (2001) 130.
[90] A. Samoson, T. Tuherm, J. Past, A. Reinhold, T. Anupold, I.
Heinmaa, Top. Curr. Chem. 246 (2005) 15.
[91] A.A. Shubin, O.B. Lapina, E. Bosch, J. Spengler, H. Knozinger, J.
Phys. Chem. B 103 (1999) 3138.
[92] J. Herzfeld, A.E. Berger, J. Chem. Phys. 73 (1980) 6021.
[93] J. Skibsted, N.C. Nielsen, H. Bildse, H.J. Jakobsen, Chem. Phys.
Lett. 188 (1992) 405.
[94] C. Jager, J. Magn. Reson. 99 (1992) 353.
[95] J. Skibsted, N.C. Nielsen, H. Bildse, H.J. Jakobsen, J. Magn.
Reson. 95 (1991) 88.
[96] A. Samoson, E. Kundla, E. Lippmaa, J. Magn. Reson. 49 (1982)
350.
[97] M.H. Cohen, F. Reif, Quadrupole Eects in Nuclear Magnetic
Resonance Studies of Solids, Academic Press Inc., New York, 1957.
[98] A. Samoson, E. Lippmaa, A. Pines, Mol. Phys. 65 (1988) 1013.
[99] A.A. Shubin, O.B. Lapina, G.M. Zhidomirov, in: IXth AMPERE
Summer School, Novosibirsk, 1987, p. 103.
[100] A.A. Shubin, O.B. Lapina, V.M. Bondareva, Chem. Phys. Lett. 302
(1999) 341.
[101] G.A. Zenkovets, G.N. Kryukova, S.V. Tsybulya, E.M. Alkaeva,
T.V. Andrushkevich, O.B. Lapina, E.B. Burgina, L.S. Dovlitova,
V.V. Malakhov, G.S. Litvak, Kinet. Catal. 41 (2000) 572.
[102] L. Frydman, J.S. Harwood, J. Am. Chem. Soc. 117 (1995) 5367.
[103] A. Medek, J.S. Harwood, L. Frydman, J. Am. Chem. Soc. 117
(1995) 12779.
[104] H.-T. Kwak, Z.H. Gan, J. Magn. Reson. 164 (2003) 369.
[105] J.-P. Amoureux, C. Fernandez, L. Frydman, Chem. Phys. Lett. 259
(1996) 347.
[106] U.G. Nielsen, H.J. Jakobsen, J. Skibsted, Solid State Nucl. Magn.
Reson. 23 (2003) 107.
[107] Z.H. Gan, J. Am. Chem. Soc. 122 (2000) 3242.
[108] Z.H. Gan, J. Chem. Phys. 114 (2001) 10845.
[109] P. Caravatti, L. Raunschweiler, R.R. Ernst, Chem. Phys. Lett. 100
(1983) 305.
[110] A. Bielecki, D.P. Burum, D.M. Rice, F.E. Karasz, Macromolecules
24 (1991) 4820.
[111] D.P. Burum, A. Bielecki, J. Magn. Reson. 94 (1991) 645.
[112] S. Steuernagel, Solid State Nucl. Magn. Reson. 11 (1998) 197.
[113] S.R. Hartmann, E.L. Hahn, Phys. Rev. B: Condens. Matter 128
(1962) 2042.
[114] D.E. Kaplan, E.L. Hahn, J. Phys. Radium 19 (1958) 821.
[115] M.J. Duer, Solid-state NMR Spectroscopy: Principles and Applications, Blackwell Science, Malden, MA, 2002.

189

[116] D.M. Follstaedt, C.P. Slichter, Phys. Rev. B: Condens. Matter 16


(1977) 21.
[117] T. Gullion, J. Schaefer, J. Magn. Reson. 81 (1989) 196.
[118] C. Brown, R. Achey, R. Fu, T. Gedris, A.E. Stiegman, J. Am. Chem.
Soc. 127 (2005) 11590.
[119] C. Hudalla, H. Eckert, R. Dupree, J. Phys. Chem. 100 (1996)
15986.
[120] C.P. Grey, W.S. Veeman, A.J. Vega, J. Chem. Phys. 98 (1993) 7711.
[121] C.P. Grey, A.J. Vega, J. Am. Chem. Soc. 117 (1995) 8232.
[122] N. Kim, C.P. Grey, Science 297 (2002) 1317.
[123] T. Gullion, Chem. Phys. Lett. 246 (1995) 325.
[124] W. Huang, A.J. Vega, T. Gullion, T. Polenova, J. Am. Chem. Soc.
129 (2007) 13027.
[125] C. Bonhomme, C. Coelho, T. Azais, L. Bonhomme-Coury, F.
Babonneau, J. Maquet, R. Thouvenot, C.R. Chim. 9 (2006) 466.
[126] H.M. Kao, K.H. Lii, Inorg. Chem. 41 (2002) 5644.
[127] L.B. Alemany, R.L. Callender, A.R. Barron, S. Steuernagel, D.
Luga, A.P.M. Kentgens, J. Phys. Chem. B 104 (2000) 11612.
[128] J.F. Stebbins, L.-S. Du, S. Kroeker, P. Neuho, D. Rice, J. Frye,
H.J. Jakobsen, Solid State Nucl. Magn. Reson. 21 (2002) 105.
[129] K.W. Feindel, K.J. Ooms, R.E. Wasylishen, Phys. Chem. Chem.
Phys. 9 (2007) 1226.
[130] K.J. Ooms, V.V. Terskikh, R.E. Wasylishen, J. Am. Chem. Soc. 129
(2007) 6704.
[131] D.L. Bryce, E.B. Bultz, Chem. Eur. J. 13 (2007) 4786.
[132] F. Jensen, Introduction to Computational Chemistry, John Wiley &
Sons, Chichester, 1999.
[133] S. Moon, S. Patchkovskii, First-principles calculations of paramagnetic NMR shifts, in: M. Kaupp, M. Buhl, V.G. Malkin (Eds.),
Calculation of EPR and NMR parameters: Theory and Applications, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2004
(Chapter 20).
[134] M. Buhl, NMR of transition metal compounds, in: M. Kaupp, M.
Buhl, V.G. Malkin (Eds.), Calculation of EPR and NMR
parameters: Theory and Applications, Wiley-VCH Verlag GmbH
& Co. KGaA, Weinheim, 2004 (Chapter 20).
[135] J. Autschbach, Struct. Bond. 112 (2004) 1.
[136] N. Pooransingh-Margolis, R. Renirie, Z. Hasan, R. Wever, A.J.
Vega, T. Polenova, J. Am. Chem. Soc. 128 (2006) 5190.
[137] N. Pooransingh, E. Pomerantseva, M. Ebel, S. Jantzen, D. Rehder,
T. Polenova, Inorg. Chem. 42 (2003) 1256.
[138] K.J. Ooms, S.E. Bolte, J.J. Smee, B. Baruah, D.C. Crans, T.
Polenova, Inorg. Chem. 46 (2007) 9285.
[139] B.A. Gee, Solid State Nucl. Magn. Reson. 20 (2006) 171.
[140] M.R. Hansen, G.K.H. Madsen, H.J. Jakobsen, J. Skibsted, J. Phys.
Chem. B 110 (2006) 5975.
[141] L. Truandier, M. Paris, C. Payen, F. Boucher, J. Phys. Chem. B
110 (2006) 21403.
[142] C.J. Pickard, F. Mauri, Phys. Rev. B: Condens. Matter 63 (2001)
245101.
[143] L. Truandier, M. Paris, F. Boucher, Phys. Rev. B: Condens. Matter
76 (2007) 035102.
[144] J. Skibsted, N.C. Nielsen, H. Bildse, H.J. Jakobsen, J. Am. Chem.
Soc. 115 (1993) 7351.
[145] J. Skibsted, C.J.H. Jacobsen, H.J. Jakobsen, Inorg. Chem. 37 (1998)
3083.
[146] U.G. Nielsen, H.J. Jakobsen, J. Skibsted, Inorg. Chem. 39 (2000)
2135.
[147] U.G. Nielsen, H.J. Jakobsen, J. Skibsted, J. Phys. Chem. B 105
(2001) 420.
[148] C.J. Fontenot, J.W. Wiench, M. Pruski, G.L. Schrader, J. Phys.
Chem. B 105 (2001) 10496.
[149] D.F. Khabibulin, A.A. Shubin, O.B. Lapina, Correlation of 51V
NMR parameters with local environment of vanadia sites, in: J.
Fraissard, O.B. Lapina (Eds.), Magnetic Resonance in Colloid and
Interface, vol. 76, NATO ASI, Kluwer, Dordrecht/Boston/London,
2002, pp. 537546.
[150] R.D. Shannon, C. Calvo, J. Solid State Chem. 6 (1973) 538.

190

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191

[151] R. Olazcuaga, J.M. Reau, G. le Flem, P. Hagenmuller, Z. Anorg.


Allg. Chem. 412 (1975) 271.
[152] A.D. Kelmers, J. Inorg. Nucl. Chem. 21 (1961) 45.
[153] A.D. Kelmers, J. Inorg. Nucl. Chem. 23 (1961) 279.
[154] R. Gopal, C. Calvo, Z. Kristallogr. 173 (1973) 67.
[155] W. Carrillo-Cabrera, H.G. von Schnering, Z. Kristallogr. 205 (1993)
271.
[156] G. Liu, J.E. Greedan, J. Solid State Chem. 110 (1994) 274.
[157] H. Kasatani, T. Umeki, H. Terauchi, J. Phys. Soc. Jpn. 61 (1992)
2309.
[158] E. Arisi, S.S.A. Palomares, F. Leccabue, B.E. Watts, G. Bocelli, F.
Calderon, G. Calestani, L. Righi, Mater. Sci. 39 (2004) 2107.
[159] Y. Oka, T. Yao, N. Yamamoto, J. Solid State Chem. 152 (2000) 486.
[160] C.E. Rice, W.R. Robinson, Acta Crystallogr. Sect. B: Struct. Sci. 32
(1976) 2232.
[161] B.C. Chakoumakos, M.M. Abraham, L.A. Boatner, J. Solid State
Chem. 109 (1994) 197.
[162] K.-J. Range, H. Meister, U. Klement, Z. Naturforsch. Teil B:
Anorg. Chem. Org. Chem. 45 (1990) 598.
[163] H. Chahboun, D. Groult, B. Raveau, Mater. Res. Bull. 23 (1988)
805.
[164] M.T. Casais, E. Gutierrez-Puebla, M.A. Monge, I. Rasines, C. RuizValero, J. Solid State Chem. 102 (1993) 261.
[165] F.C. Hawthorne, C. Calvo, J. Solid State Chem. 26 (1978) 345.
[166] J. Huang, A.W. Sleight, Mater. Res. Bull. 27 (1992) 581.
[167] R.D. Shannon, C. Calvo, Can. J. Chem. 51 (1973) 265.
[168] F. Marumo, M. Isolbe, S. Iwai, Y. Kondo, Acta Crystallogr. Sect. B:
Struct. Sci. 30 (1974) 1628.
[169] H.T. Evans, Z. Kristallogr. 114 (1960) 257.
[170] M. Ganne, Y. Piard, M. Tournoux, Can. J. Chem. 52 (1974) 3539.
[171] F.C. Hawthorne, C. Calvo, J. Solid State Chem. 22 (1977) 157.
[172] T. Yao, Y. Oka, N. Yamamoto, Inorg. Chim. Acta 238 (1995) 165.
[173] N.N. Hok, C. Calvo, Can. J. Chem. 50 (1972) 3619.
[174] G.D. Andreetti, G. Calestani, A. Montenero, Z. Kristallogr. 168
(1984) 53.
[175] K. Mocala, J. Ziolkowski, J. Solid State Chem. 69 (1987) 299.
[176] O.G. Karpov, M.A. Simonov, T.I. Krasnenko, O.A. Zabara,
Kristallograya 34 (1989) 1392 (in Russian).
[177] N. Das, H. Eckert, H. Hu, I.E. Wachs, J.F. Walzer, F.J. Feher, J.
Phys. Chem. 97 (1993) 8240.
[178] Y. Oka, S. Fumihiko, Y. Takeshi, Y. Naoichi, J. Solid State Chem.
134 (1997) 52.
[179] F.B. Abdelouahab, R. Olier, N. Guilhaume, F. Lefebre, J.C. Volta,
J. Catal. 134 (1992) 151.
[180] O.B. Lapina, D.F. Khabibulin, A.A. Shubin, V.M. Bondareva, J.
Mol. Catal. A: Chem. 162 (2000) 381.
[181] E.V. Sokolova, M.A. Simonov, U.V. Karyakin, A.V. Zavolskaya,
Kristallograya 28 (1983) 862 (in Russian).
[182] H.T.J. Evance, Inorg. Chem. 5 (1966) 967.
[183] Y. Shan, S.D. Huang, J. Chem. Crystallogr. 29 (1999) 93.
[184] O.W. Howarth, R.E. Richards, J. Chem. Soc. (1965) 864.
[185] H. Eckert, I.E. Wachs, J. Phys. Chem. 93 (1989) 6796.
[186] S. Hayakawa, T. Yoko, S. Sakka, Bull. Chem. Soc. Jpn. 66 (1993)
3393.
[187] U.G. Nielsen, A. Boisen, M. Brorson, C.J.H. Jacobsen, H.J.
Jakobsen, J. Skibsted, Inorg. Chem. 41 (2002) 6432.
[188] R. Cousin, D. Courcot, E. Abi-Aad, S. Capelle, J.P. Amoureux, M.
Dourdin, M. Guelton, A. Aboukais, Colloids Surf. A 158 (1999) 43.
[189] F. Delmaire, M. Rigole, E.A. Zhilinskaya, A. Aboukais, R. Hubaut,
G. Mairesse, Phys. Chem. Chem. Phys. 2 (2000) 4477.
[190] P.R. Bodart, Y. Dumazy, J.P. Amoureux, C. Fernandez, Magn.
Reson. Chem. 37 (1999) 223.
[191] O.B. Lapina, V.M. Mastikhin, A.V. Nosov, T. Beutel, H. Knozinger, Catal. Lett. 13 (1992) 203.
[192] P.E. Stallworth, X. Guo, E. Tatham, S.G. Greenbaum, M. Arrabito,
S. Bodoarado, N. Penazzi, Solid State Ionics 170 (2004) 181.
[193] D.F. Khabibulin, K.V. Romanenko, Z.H. Gan, E. Arkhipova, M.
Zuev, O.B. Lapina, Magn. Reson. Chem. 45 (2007) 962.

[194] S. Hayakava, Y. Toshinobu, S. Sumio, J. Solid State Chem. 112


(1994) 329.
[195] A.A. Shubin, O.B. Lapina, in: V Russian Conference on Structure
and Dynamics of Molecular Systems, vol. 1, Mariisky State
University, Ioshkar Ola, 1998, p. 3641 (in Russian).
[196] C. Fernandez, P. Bodart, J.-P. Amoureux, Solid State Nucl. Magn.
Reson. 3 (1994) 79.
[197] O. Durupthy, N. Steunou, T. Coradin, J. Maquet, C. Bonhomme, J.
Livage, J. Mater. Chem. 15 (2005) 1090.
[198] R. Siegel, N. Dupre, M. Quarton, J. Hirschinger, Magn. Reson.
Chem. 42 (2004) 1022.
[199] O.B. Lapina, V.V. Terskikh, A.A. Shubin, K.M. Eriksen, R.
Fehrmann, Colloids Surf. A 158 (1999) 255.
[200] W. Huang, L. Todaro, L.C. Francesconi, T. Polenova, J. Am. Chem.
Soc. 125 (2003) 5928.
[201] W. Huang, L. Todaro, G.P.A. Yap, R. Beer, L.C. Francesconi, T.
Polenova, J. Am. Chem. Soc. 126 (2004) 11564.
[202] O. Durupthy, M. Jaber, N. Steunou, J. Maquet, G.T. Chandrappa,
J. Livage, Chem. Mater. 17 (2005) 6395.
[203] D. Franklin, I.E. Wachs, H. Eckert, D.A. Jeerson, J. Solid State
Chem. 90 (1991) 194.
[204] B. Alonso, C. Sanchez, J. Mater. Chem. 10 (2000) 377.
[205] K. Ukei, H. Suzuki, T. Shishido, T. Fukuda, Acta Crystallogr. Sect.
C: Cryst. Struct. Commun. 50 (1994) 655.
[206] T. Shishido, H. Suzuki, K. Ukei, T. Hibiya, T. Fukuda, J. Alloys
Compd. 234 (1996) 256.
[207] A. Castro, E. Aguado, J.M. Rojo, P. Herrero, R. Enjalbert, J. Galy,
Mater. Res. Bull. 33 (1998) 31.
[208] A. Kahn-Harari, L. Mazerolles, D. Michel, F. Robert, J. Solid State
Chem. 116 (1995) 103.
[209] M.A. Subramanian, J.C. Calabrese, Mater. Res. Bull. 28 (1993)
523.
[210] R. Hsu, E.N. Maslen, D.D. Boula, N. Ishizawa, Acta Crystallogr.
Sect. B: Struct. Sci. 53 (1997) 420.
[211] A.W. Hewat, J. Phys. C: Sol. St. Phys. 6 (1973) 2559.
[212] V.A. Shuvaeva, M.Y. Antipin, Kristallograya 40 (1995) 511 (in
Russian).
[213] T.S. Ercit, P. Cerny, Can. Mineral. 26 (1988) 899.
[214] J. Galy, O. Lindqvist, J. Solid State Chem. 27 (1979) 279.
[215] K.J.D. MacKenzie, M.E. Smith, Multinuclear Solid State NMR of
Inorganic Materials, Pergamon Elsevier, Oxford, 2002.
[216] Y. Watanabe, T. Sota, K. Suzuki, N. Iyii, K. Kitamura, S. Kimura,
J. Phys.: Condens. Matter 7 (1995) 3627.
[217] M.D. Meadows, K.A. Smith, R.A. Kinsey, M. Rothgeb, R.P.
Skarjune, E. Oldeld, Proc. Natl. Acad. Sci. USA 79 (1982) 1351.
[218] H. Weitzel, H. Schroecke, Z. Kristallogr. 152 (1980) 69.
[219] S. Tsunekawa, T. Kamiyama, K. Sasaki, H. Asano, T. Fukuda, Acta
Crystallogr. Sect. A: Found. Crystallogr. 49 (1993) 595.
[220] N.A. Godina, T.I. Panova, E.K. Keller, Izvestiya AN SSSR Inorg.
Mater. 5 (1969) 1974 (in Russian).
[221] K.-J. Range, M. Wildenauer, A.M. Heyns, Angew. Chem. Int. Ed.
Engl. 27 (1988) 969.
[222] A. Flambard, L. Montagne, L. Delevoye, S. Steuernagel, Solid State
Nucl. Mag. Reson. 32 (2007) 34.
[223] J.P. Jesson, The paramagnetic shifts, in: G.N. La Mar, W.D.
HorrocksJr., R.H. Holm (Eds.), NMR of Paramagnetic Molecules:
Principles and Applications, Academic Press, New York and
London, 1973, pp. 152.
[224] I. Bertini, C. Luchinat, Coord. Chem. Rev. 150 (1996) 1.
[225] I. Bertini, C. Luchinat, G. Parigi, Prog. Nucl. Magn. Reson.
Spectrosc. 40 (2002) 249.
[226] T. Moriya, Prog. Theor. Phys. 16 (1956) 23.
[227] V. Jaccarino, Nuclear resonance in antiferromagnets, in: G.T. Rado,
H. Suhl (Eds.), Magnetism, vol. 2, Academic Press, New York, 1965,
p. 307.
[228] M.I. Kurkin, E.A. Turov, NMR in Magnetically Ordered Materials
and its Applications, Nauka, Moscow, 1990 (in Russian).
[229] P. Lemmens, P. Millet, Lect. Notes Phys. 645 (2004) 433.

O.B. Lapina et al. / Progress in Nuclear Magnetic Resonance Spectroscopy 53 (2008) 128191
[230] J. Stohr, H.C. Siegmann, Magnetism from fundamentals to nanoscale dynamics, Springer Series in Solid-State Sciences, vol. 152,
Springer-Verlag, Berlin, Heidelberg, 2006.
[231] J. Li, M.E. Lashier, G.L. Schrader, B.C. Gerstein, Appl. Catal. 73
(1991) 83.
[232] A. Tuel, L. Canesson, J.C. Volta, Colloids Surf. A 158 (1999) 97.
[233] R.N. Pletnev, V.N. Lisson, V.A. Gubanov, Fizika Tverdogo Tela 15
(1973) 558 (in Russian).
[234] R.N. Pletnev, A.A. Sidorov, V.N. Lisson, A.K. Chirkov, Dokladi
Akademii Nauk SSSR 236 (1977) 1159 (in Russian).
[235] R. Schowalter, Phys. Status Solidi B 48 (1971) 743.
[236] U. Ranon, Phys. Lett. A 28 (1968) 228.
[237] A.A. Fotiev, V.L. Volkov, V.K. Kapustkin, Bronses of Vanadium
Oxides, Nauka, Moskva, 1978 (in Russian).
[238] D. Tinet, J.J. Fripiat, Rev. Chim. Mineral. 19 (1982) 612.
[239] K. Nakamura, D. Nishioka, Y. Michihiro, M. Vijayakumar, S.
Selvasekarapandian, T. Kanashiro, Solid State Ionics 177 (2006) 129.
[240] N.I. Lazukova, V.A. Gubanov, Zh. Struk. Khim. 18 (1977) 10.
[241] F.-N. Shi, F.A.A. Paz, J. Rocha, J. Klinowski, T. Trindade, Eur. J.
Inorg. Chem. 2004 (2004) 3031.
[242] U.G. Nielsen, J. Skibsted, H.J. Jakobsen, Chem. Phys. Lett. 356
(2002) 73.
[243] I. Karatchevtseva, Z. Zhang, J. Hanna, V. Luca, Chem. Mater. 18
(2006) 4908.
[244] F. Carn, N. Steunou, J. Livage, A. Colin, R. Backov, Chem. Mater.
17 (2005) 644.
[245] N. Dupre, J. Gaubicher, D. Guyomard, C.P. Grey, Chem. Mater. 16
(2004) 2725.
[246] M. Dubarry, J. Gaubicher, D. Guyomard, N. Steunou, J. Livage, N.
Dupre, C.P. Grey, Chem. Mater. 18 (2006) 629.
[247] R. Michalaka, H.A.H. Mohammad, Solid State Nucl. Magn. Reson.
26 (2004) 187.
[248] L.G. Pinaeva, O.B. Lapina, V.M. Mastikhin, A.V. Nosov, B.S.
Balzhinimaev, J. Mol. Catal. 88 (1994) 311.
[249] D. Courcot, B. Grzybowska, Y. Barbaux, M. Rigole, A. Ponchel, M.
Guelton, J. Chem. Soc., Faraday Trans. 92 (1996) 1609.

191

[250] L. Lietti, P. Forzatti, G. Ramis, G. Busca, F. Bregani, Appl. Catal.


B 3 (1993) 13.
[251] G.T. Went, L.J. Leu, S.J. Lombardo, A.T. Bell, J. Phys. Chem. 96
(1992) 2235.
[252] V.M. Mastikhin, I.L. Mudrakovsky, A.V. Nosov, Prog. Nucl.
Magn. Reson. Spectrosc. 23 (1991) 259.
[253] G. Deo, I.E. Wachs, J. Phys. Chem. 95 (1991) 5889.
[254] C.J. Fontenot, J.W. Wiench, M. Pruski, G.L. Schrader, J. Phys.
Chem. B 104 (2000) 11622.
[255] C.J. Fontenot, J.W. Wiench, G.L. Schrader, M. Pruski, J. Am.
Chem. Soc. 124 (2002) 8435.
[256] D.C. Sayele, C.R.A. Catlow, M.-A. Perrin, P. Nortier, J. Phys.
Chem. 100 (1996) 8940.
[257] U.G. Nielsen, N.U. Topse, M. Brorson, J. Skibsted, H.J. Jakobsen,
J. Am. Chem. Soc. 126 (2004) 4926.
[258] O.B. Lapina, A.A. Shubin, A. Nosov, E. Bosch, J. Spengler, H.
Knozinger, J. Phys. Chem. 103 (1999) 7599.
[259] C. Martin, D. Klissurski, J. Rochac, V. Rives, Phys. Chem. Chem.
Phys. 2 (2000) 1543.
[260] V.V. Terskikh, O.B. Lapina, V.M. Bondareva, Phys. Chem. Chem.
Phys. 2 (2000) 2441.
[261] R. Blinc, B. Zalar, A. Gegorovic, R. Pirc, M.D. Glinchuk,
Ferroelectrics 240 (2000) 1473.
[262] R. Pirc, R. Blinc, Phys. Rev. B: Condens. Matter 60 (1999) 13470.
[263] V.V. Laguta, M.D. Glinchuk, I.P. Bykov, R. Blinc, B. Zalar, Phys.
Rev. B: Condens. Matter 69 (2004) 54103/1.
[264] A.Y.H. Lo, T.E. Bitterwolf, C.L.B. Macdonald, R.W. Schurko, J.
Phys. Chem. A 109 (2005) 7073.
[265] S. Rankel, B. Zalar, V.V. Laguta, R. Blinc, J. Toulouse, Ferroelectrics 314 (2005) 165.
[266] K. Sato, S. Takeda, S. Fukuda, T. Minamisono, M. Tanigaki, T.
Miyake, Y. Maruyama, K. Matsuta, M. Fukuda, Y. Nojiri, Z.
Naturforsch. A: Phys. Sci. 53 (1998) 549.
[267] O.B. Lapina, V.M. Mastikhin, L.G. Simonova, Y.O. Bulgakova, J.
Mol. Catal. 69 (1991) 61.
[268] J. Klinowski, Prog. Nucl. Magn. Reson. Spectrosc. 16 (1984) 237.

Vous aimerez peut-être aussi