Você está na página 1de 12

Solid State Nuclear Magnetic Resonance ∎ (∎∎∎∎) ∎∎∎–∎∎∎

Contents lists available at ScienceDirect

Solid State Nuclear Magnetic Resonance


journal homepage: www.elsevier.com/locate/ssnmr

Trends

NMR crystallography: Applications to inorganic materials


Charlotte Martineau
Tectospin, Institut Lavoisier de Versailles, UMR CNRS 8180, Université de Versailles St-Quentin en Yvelines, 45, avenue des Etats-Unis,
78035 Versailles cedex, France

art ic l e i nf o a b s t r a c t

Article history: Current developments of NMR crystallography as well as some recent applications to diamagnetic
Received 5 June 2014 inorganic solids are presented. First, we illustrate how solid-state NMR data can be used in combination
Received in revised form with diffraction data for the determination of the periodic part of the crystal structures, from the space
3 July 2014
group selection, to the structure determination over the refinement and validation processes. As ss-NMR,
contrary to diffraction (powder and single-crystal), is not restricted to periodic boundary conditions, ss-
Keywords: NMR data can be used to further complete the structural description of materials, including studies of
NMR crystallography local order/disorder, etc. This illustrated through examples, which are shown and discussed in the second
Inorganic compounds part of this review.
Powder diffraction
& 2014 Elsevier Inc. All rights reserved.
Structure solution
Structure refinement
Local order

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2. Periodic model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.1. Determination of the unit cell dimensions and space group. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.2. Building a structural model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2.1. From NMR data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2.2. From powder diffraction and NMR data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.3. Structure refinement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
3. Beyond periodicity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.1. Cation distribution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2. Anion distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.3. Linker distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.4. Correlated disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.5. Defects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.6. Glasses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4. Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

1. Introduction to the study of molecular and crystalline structure and properties,


with far-reaching applications in mineralogy, chemistry, physics,
According to the International Union of Crystallography (IUCr), mathematics, biology and materials science.” [1]. 2014 is the Inter-
crystallography [from Greek κρυσταλλος (krustallos) ice, crys- national Year of Crystallography, which celebrates “the 100th
tal þ γραφειν (graphein) to write] “is the branch of science devoted anniversary of X-ray diffraction as well as the 400th anniversary
of Kepler's observation in 1611 of the symmetrical form of ice
crystals, which began the wider study of the role of symmetry in
E-mail address: charlotte.martineau@uvsq.fr matter”. If X-ray diffraction is by far the most powerful and widely

http://dx.doi.org/10.1016/j.ssnmr.2014.07.001
0926-2040/& 2014 Elsevier Inc. All rights reserved.

Please cite this article as: C. Martineau, Solid State Nucl. Magn. Reson. (2014), http://dx.doi.org/10.1016/j.ssnmr.2014.07.001i
2 C. Martineau / Solid State Nuclear Magnetic Resonance ∎ (∎∎∎∎) ∎∎∎–∎∎∎

used technique to unravel the crystalline structure of solids, it is offer. It is also particularly important for fully disordered systems
not the only available tool we have at hand as there are many like glasses for which the information available on powder
other spectroscopic techniques which are complementary with diffraction patterns are very limited.
powder diffraction, like electron microscopy, infra-red (IR), X-ray The major difference between using powder diffraction and
absorption near edge structure (XANES), extended X-ray absorp- ss-NMR – in particular for inorganic solids which may contain a
tion fine structure (EXAFS) or nuclear magnetic resonance (NMR) large number of different elements - comes from the fact that
spectroscopy. diffraction requires a rather limited number of experimental data
The study of the structure of the periodic part of crystalline (usually a single X-ray powder diffraction measurement), while an
solids in fact requires a rather limited amount of information, almost infinite number of NMR measurements can be imagined,
which are (i) the unit cell dimensions, (ii) the space group, (iii) the designed and performed, each of them yielding different structural
nature of the atoms and (iv) their relative position in the unit cell. data. This probably explains why there is not yet a very general
One way of accessing these pieces of information is to submit the strategy for NMR Crystallography (as there exists in diffraction),
solids to a beam of X-ray. The Bragg's law and structure factors but rather there are many different trends explored, which are
indeed relate the diffraction peak positions and intensity to the very solid- (or class or solid) dependant. NMR Crystallography is
electron density and the atomic coordinates. These principles are an active field of research, as attested by the special issue on the
nowadays easily and widely applied to single-crystals, for which subject published last year [11,12]. A few of the strategies devel-
the diffraction peaks can be measured in the three directions of oped recently for inorganic materials are covered and discussed
space, and the structure reconstructed from these data. Single- here. This review is by no mean exhaustive, but reports some
crystal diffraction has therefore provided a huge number of examples in the fields of oxides [13–26], fluorides [27–33], porous
successes, in a wide range of solids, including inorganic solids. solids like Al- [34–40], Ti- [41,42], Ga- [43,44], or V- [45,46] based
There is however one critical limitation of this technique, as it phosphates or oxonitrophosphate [47], zeolites, metal-organic
requires that crystals of large enough size, stability and quality are frameworks (MOFs) based on metals like 27Al [48–53], Ga [54],
45
grown. Although recent development of microdiffraction [2] setup Sc [55], 43Ca [56,57], 25Mg, 67Zn [58], Zr [59,60], 17O [61] etc. One
has considerably reduced the minimum required size of the crystal goal of this article is to show how the strengths of ss-NMR data are
down to less than a few cubic microns, there is still an incredibly really decupled when they are combined with powder diffraction,
large number of materials for which it is not possible to grow such electron diffraction, as well as other spectroscopic techniques and
single-crystals. In this case, the structure determination has to be calculations of NMR observables.
carried out from X-ray diffraction data collected on polycrystalline
powders. This poses a serious problem, as crucial information, in
particular the phases of the diffraction peaks, is lost when going 2. Periodic model
from a 3D measure on a single-crystal to a 2D projection over a
powder, an ensemble of tiny single-crystals, each having its own As mentioned earlier, there are four mains step to determine
orientation. For this purpose, a series of methods and algorithms the periodic structure of a compound [62] (Fig. 1) (i) finding the
(direct methods [3], Patterson method [4], charge-flipping [5], real unit cell dimensions, and the space group, (ii) finding the nature of
space methods [6] with simulated annealing [7] and its derivative the atoms and their approximate relative positions, that is, build-
parallel tempering [8] algorithms, etc) and associated softwares ing a structural ‘model’, (iii) refining this model, i.e., finding atomic
have been developed – and are still being developed – to try to coordinates that match the best an experimental function (usually
retrieve the missing information from powder diffraction patterns. the powder diffractogram). If all these steps can be performed, in
However, for ab initio structure determination, i.e., when there is principle, from powder diffraction data only, there are numerous
no prior knowledge about the structure, this process can become cases where, due to peak overlap or other ambiguities on the
highly time-consuming, and even often unsuccessful at all for very diagram, (i) the solution might not be unique, (ii) no solution is
complex systems. found because the solution space to be searched is too vast or its
It turns out that solid-state NMR (ssNMR) spectra contain landscape contains too many local minima, preventing conver-
relevant data for structure determination and, these pieces of gence of the calculations in a reasonable amount of time.
information are, interestingly, highly complementary to the dif- As illustrated in Fig. 1, ss-NMR data can be used along with the
fraction data. NMR spectroscopy is indeed sensitive to the nature diffraction data to increase the chances of success of each of these
of the atom through the Larmor frequency, and to the local steps, and sometimes reduce the time necessary to perform these
environment of each atom, while diffraction is based on periodi- steps. If initially the NMR data were used only as a validation tool
city and relies on electron density to distinguish the atoms. for a structure independently determined from powder diffraction
Moreover, the progress in NMR methods and techniques (high data, it was recently shown that NMR data could be used prior to
field, fast magic-angle spinning (MAS), pulse sequences for high- the structure search, to guide and make more efficient the process.
resolution, sensitivity enhancement techniques for faster acquisi- Some of the most recent examples illustrating each stage are
tion etc) has given access to the resolution and visibility of reported and discussed in the next paragraphs.
structural features, even for complex materials. More recently,
the developments of first-principles calculations to periodic com- 2.1. Determination of the unit cell dimensions and space group
pounds have offered another opportunity to link the structural
data to NMR observables. The term NMR Crystallography [9] has Determination of the unit cell dimension and space group is
been proposed a decade ago, with the initial idea to perform traditionally done by indexing the powder diffraction pattern.
structure determination from NMR data only, but now it encom- Determination of the unit cell dimensions is still very difficult – if
passes all studies in which the NMR data are used along with the not possible – to perform from ss-NMR data, and this is the major
diffraction, first-principles calculations of other spectroscopic item that is left to diffraction (X-ray or electrons) only. It can seem
techniques to perform the structure determination, and go beyond a rather trivial step, but this is in fact the most important one, as
what diffraction proposes, i.e., beyond strict periodicity. The without a unit cell, the next steps of structure determination could
description of both the periodic and non-periodic parts of solids not be achieved, as they all somehow involved calculations within
has been defined by Mackay [10] as Generalized Crystallography, a ‘box’. The space group can be found by searching for systematic
and this is what the combination of NMR and diffraction has to extinctions of (hkl) reflections. However, (i) only 120 extinction

Please cite this article as: C. Martineau, Solid State Nucl. Magn. Reson. (2014), http://dx.doi.org/10.1016/j.ssnmr.2014.07.001i
C. Martineau / Solid State Nuclear Magnetic Resonance ∎ (∎∎∎∎) ∎∎∎–∎∎∎ 3

Fig. 1. Schematic representation of the main steps for ab initio structure resolution. In blue and red the information or methods used with powder diffraction and ss-NMR
data, respectively. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

symbols are univocally related to one space group [63], (ii) strong 2.2. Building a structural model.
peak overlap may generate ambiguities. Therefore, if the Laue class
can usually be determined, it is very often that line indexing does That is indisputably the most difficult and hazardous step in
not lead to a unique solution. the whole process, where failure is very (very) often encountered.
The role of ss-NMR at this stage is thus to provide comple- This is therefore the step where ss-NMR has probably the most
mentary data to select the proper space group among the list of impact. At this stage, the goal is to build an approximate structural
space groups provided by the diffraction data. NMR data can give model (number, nature and connectivity of the cation polyhedra,
access to the ‘crystallochemical formula’. This formula gathers the presence of organic moieties, guest molecules, etc.). When no iso-
nature, number and relative intensity of all the atoms of the structural compounds or compounds with a closely related struc-
asymmetric unit, the asymmetric unit being the smallest unit that ture exist, this has to be done ab initio, i.e. blindly. A number of
will be repeated through symmetry operations to create the 3D methods have been developed over the years to perform this
structure. A space group, to be compatible with a structure, must ab initio structure determination, as already mentioned in the
contain enough ‘Wyckoff positions’ to accommodate all the introduction section. Although these algorithms have been
measured atoms. The principles of space group selection by NMR improved and are now more and more efficient, there are many
were exposed by Taulelle some ten years ago [64]. In short, it reasons that can make the structure determination difficult from
makes sure that the space group contains Wyckoff positions with PXRD data only, among which the problem of peak overlaps
appropriate relative multiplicity, and that if there are atoms leading to unreliable extracted line intensities, the size of the unit
located on special positions with no degree of freedom, there are cell (too large), the ratio between the number of atomic para-
enough of these positions in the chosen space group to accom- meters and the number of observed reflections (which has to be
modate all these atoms. In principle it is a useful concept for low enough), extreme atomic contrast (either too large or too low),
inorganic solids, as virtually all the space groups can be considered possible disorder, poor crystallinity, computation time, etc.
(contrary to organic solids or proteins, most of which crystallize in The role of ss-NMR can be of two types: (i) construct an initial
monoclinic space groups). However, in practice, this method – i.e., model from the NMR data only, (ii) provide constraints (distances,
space group selection by NMR only – mostly works for compounds topology), so as to drive more efficiently the search for the
with highly symmetrical space groups, and containing atoms structural model from the powder diffraction data. Both are
sitting on special positions with no degree of freedom. Therefore, detailed below.
the most efficient use of ss-NMR is, through the determination of
the number of inequivalent sites, to sort out among the space
groups provided by the diffraction data [15,29,43,58,65–69]. 2.2.1. From NMR data
In particular, it can ascertain or dismiss the presence of a center It is well known that homonuclear and heteronuclear inter-
of symmetry, which sometimes is ambiguous from diffraction nuclear distances can be measured from ss-NMR data, by measur-
data only. ing e.g. diffusion rates, build-up curves (DQ, REDOR, etc.), or

Please cite this article as: C. Martineau, Solid State Nucl. Magn. Reson. (2014), http://dx.doi.org/10.1016/j.ssnmr.2014.07.001i
4 C. Martineau / Solid State Nuclear Magnetic Resonance ∎ (∎∎∎∎) ∎∎∎–∎∎∎

Fig. 2. (a) 23Na 2D NMR experiment recorded under DOR conditions. (b) Spin diffusion curves constructed from a series of 2D 23
Na NMR measurements in a sodium
diphosphate compound [72]. Reproduced by permission of the Royal Society of Chemistry.

Please cite this article as: C. Martineau, Solid State Nucl. Magn. Reson. (2014), http://dx.doi.org/10.1016/j.ssnmr.2014.07.001i
C. Martineau / Solid State Nuclear Magnetic Resonance ∎ (∎∎∎∎) ∎∎∎–∎∎∎ 5

through the measurement of some NMR parameters (chemical


shift, quadrupolar parameters, etc). This property has initially been
developed for NMR Crystallography of small organic molecules,
with the measurement of 1H–1H spin diffusion rates [70].
It provides a distance matrix, from which the relative positions
of the molecules in the unit cell can be worked out.
Recently, Bryce et al. showed the possibility to extend the
method to quadrupolar nuclei for inorganic materials, thanks to
the use of double rotation (DOR) techniques, which provides the
necessary high-resolution in the 23Na (nuclear spin equals to 3/2)
2D NMR spectra (Fig. 2) [71]. Applied to sodium diphosphates
Na4P2O7 and Na3HP2O7  H2O, spin diffusion curves were mea-
sured, modeled and the dipolar coupling values – thus distances –
were extracted. These values were then used to validate the
assignment of the NMR resonances.
For pure-silica zeolites, Brouwer et al. proposed a method to
build a structural model starting with initial knowledge of (i) the
unit cell dimension and space group as determined from powder
diffraction and (ii) a homonuclear distance matrix [72–78]. Zeo-
lites have a structures based on the assembly of corner-sharing
SiO4 tetrahedra (T-sites), forming porous framework, in which are Mcorr Pa {Pb, Pc}* Pd Pe {Pf, Pg}*
located the templating agents before calcination (cations or
organic moieties). They are notoriously difficult to obtain as {Ala, 0 1 1 0 1 1 1
single-crystal, and the structure determination of these com-
pounds is still a challenge by powder X-ray diffraction only. Great Alb} 0 1 1 0 1 1 1
success was obtained with the additional use of electron micro-
scopy with the charge-flipping algorithm [79–81]. NMR provides Alc 1 1 1 1 1 0 0
another source of information.
The homonuclear distance matrix can indeed be constructed Ald 1 1 1 1 0 1 1
through the measurement of 29Si double-quanta (DQ) build-up
curves [82–84]. The low natural abundance of 29Si nucleus Ale 1 1 1 0 1 1 1
(nuclear spin equals to 1/2, natural abundance of 4%) is favorable
to this method as the assumption that DQ curves are obtained for Alf 0 1 1 1 0 1 1
isolated 29Si–29Si spin pair is valid, a requisite to go from the
measured build-up curves to the actual Si–Si internuclear dis-
Fig. 3. Top: 27Al–31P 2D correlation NMR spectrum of an aluminophosphate and
tances with analytical expression. Depending on the recoupling
bottom: corresponding graph matrix [88].
method, the 29Si–29Si DQ-build-up curve provide access to Si–Si
distances up to about 8 Å between pairs of atoms.
As mentioned above, with the knowledge of the unit cell be abstracted as the nodes of the graph, while the bridging oxygen
dimension and possibly space group (previously extracted from atoms (or the organic linkers for MOFs) represent the edges of this
powder diffraction data), an algorithm was developed which graph. In NMR, the resonances represent the number of nodes
allows a step-by-step construction of a candidate structure [74], (independent of the nature of the atom and of its coordination
which is in a later stage optimized (by energy minimization and/or number), while the correlation peaks on a 2D NMR spectrum
refinement from the XRD data, when the materials are crystalline indicate the number of the graph edges, and which nodes are
enough). Recent improvements of this method have been pro- connected through these vertices. Brouwer et al. recently showed
posed, which can yield a structure model even when the crystal- how such graph theory could be used to reconstruct a silicate
lographic space group is unknown (i.e., only unit cell dimensions network from a limited number of experimental data [88]: one
are known a priori). The principles were illustrated for a hypothe- powder diffraction pattern, that provides the unit cell dimensions
tical 2D zeolite structure, and have then been successfully applied (this is still a requirement for all the structure determination
to the structural determination of molecularly ordered but non- strategies), and a single 2D 29Si–29Si NMR spectrum (assuming all
crystalline silicate framework [85]. NMR lines are resolved), that provides the number of inequivalent
Si sites along with their relative multiplicity, the number of
equivalent positions in the unit cell (Z) as well as the Si–Si
2.2.1.1. Graph theory. For compounds which form extended 2- or connectivities. Knowledge about the space group is not manda-
3D frameworks (as is the case for most zeolites, aluminophosphates tory. One strong limitation is that the graph generation is sub-
or metal-organic frameworks (MOFs)), the concept of graph theory mitted to combinatorial explosion of computation time, and at
(no consideration of distance, one only considers the link between present can only be applied, in a reasonable amount of time, to
the atoms) can be used to build a structural model. A graph is made simple frameworks (i.e., below 4 vertices). They mention that the
of nodes (also called vertices) connected to each other through approach can be extended to heteronuclear systems like
linkers. The use of graph to describe the structure of solids is not new aluminophosphates.
and for example the software TOPOS [86] is a very handy tool to
identify isomorphic compounds (i.e., compounds having different
structures but similar topology). 2.2.2. From powder diffraction and NMR data
It turns out that both the frameworks of porous solids and Real space methods are the most recent and efficient methods
ss-NMR correlation spectra can be defined as graphs (Fig. 3 [87]). for structure solution from powder diffraction, and probably the
The cation polyhedra (or inorganic clusters for MOFs) can indeed methods for which NMR can have the most beneficial impact. Real

Please cite this article as: C. Martineau, Solid State Nucl. Magn. Reson. (2014), http://dx.doi.org/10.1016/j.ssnmr.2014.07.001i
6 C. Martineau / Solid State Nuclear Magnetic Resonance ∎ (∎∎∎∎) ∎∎∎–∎∎∎

space methods use algorithms like simulated annealing, parallel independent parameters is then reduced from three times N
tempering or charge flipping and require as a priori knowledge a coordinates for each of the N atoms down to three atomic
unit cell, if possible a space group (not mandatory, e.g., the charge coordinate and three angles for the ‘object’. In addition to having
flipping algorithm can work in P1), and the approximate number the exact number of species in the unit cell, knowledge about the
and nature of atoms in the unit cell. They then randomly generate coordination number already represents a great step towards
candidate structures until convergence to a given cost function success.
(total energy, powder diffraction diagram) is reached. In these Using these pieces of information, structural model can be
Monte-Carlo based methods, computation time quickly becomes a determined, even for compounds that show difficult powder
strong limitation, even for efficient (parallel) computers we have diffractogram (poor crystallinity, strong absorption), as illustrated,
nowadays. For a unit cell containing a number N of inequivalent for example, for the structures of a few phosphonates that were
sites, the convergence rate indeed grows as exp(N2). Therefore, recently reported by Laurencin et al.. The synchrotron powder
one limitation is the number of independent parameters (atomic diffraction showed strong profile anisotropy due to the layered
coordinates for each ‘object’ in the unit cell), as often mentioned character of these compounds, as well as strong baseline distor-
by the authors of the programs: the limit of ESPOIR [89] seems to tion. Nonetheless, the unit cell and space groups could be
be close to 15–30 independent atoms maximum or 6 independent determined by indexation. 1H, 13C, 31P and 43Ca NMR solid-state
objects, the FOX [90] and PSSP [91] software capabilities have been NMR data further provided the number of inequivalent Ca sites,
tested for only 6–20 degrees of freedom. Finally, the software phosphonate groups and their protonation states, as well as the
FOCUS [92], specifically designed to perform structure search on number of water molecules in the unit cell (Fig. 4). These species
materials like zeolite or aluminophosphate requires time to expect were used as input independent ‘objects’ in the software FOX [92],
relevant structural models ranging from a few days to several which converged to a sensible solution for all four layered samples
months. studied Ca(C6H5–PO3H)2, Ca(C6H5–PO3)  2H2O, Ca(C4H9–PO3H)2
The major role of NMR combined with these methods is and Ca(C4H9–PO3)  H2O. As discussed in their paper, the structural
therefore to lower this critical number of parameters down to a models were then refined at the DFT level and validated by first-
level that becomes computable again. principles calculations of the NMR parameters, which were in very
good agreement with the experimental data [93].
2.2.2.1. Knowledge about the unit cell content. The first obvious
constraint to extract from NMR data are the exact number of 2.2.2.2. Adjacency matrix. A second way to reduce the number of
inequivalent atoms in the asymmetric unit, as well as the independent parameters is to use larger units of connected
coordination number of the atoms, which can usually be polyhedra (instead of using the polyhedra as isolated species).
straightforwardly determined from NMR parameters like This concept was recently proposed for aluminophosphates
chemical shifts or quadrupolar parameters. This latter allows (AlPOs) [98]. Under hydrothermal synthesis conditions, the struc-
considering the full polyhedron (or molecule when organic ture of these solids is made from strict alternation of aluminum
moieties are present) as an ‘object’, rather than considering the polyhedra and phosphorus tetrahedra connected through bridging
cations and linker atoms each as isolated species. The number of oxygen atoms. Compared to zeolites, AlPOs present a much larger

40 35 30 25 20 15 10 100 80 60 40 20 0 -20 -40 -60 -80-100


43
13 Ca chemical shift (ppm)
C chemical shift (ppm)

Fig. 4. Schematic representation of combining diffraction data (determination of unit cell dimension and space group) and solid-state NMR (determination of the number
and nature of the species). The structure search is then performed with a direct space software (FOX in that case), with the powder diffraction diagram as experimental cost
function. The example shows the data for Ca(BuPO3).

Please cite this article as: C. Martineau, Solid State Nucl. Magn. Reson. (2014), http://dx.doi.org/10.1016/j.ssnmr.2014.07.001i
C. Martineau / Solid State Nuclear Magnetic Resonance ∎ (∎∎∎∎) ∎∎∎–∎∎∎ 7

variety of topologies because of the ability of the metal cations (in It must be noted that, for this step, the methods reported above
particular Al3 þ , Ga3 þ …) to adopt various coordination states mostly (only) makes use of the NMR information regarding the
within the lattice: from tetrahedral to octahedral through penta- cation(s), rarely regarding the linker atoms (in particular oxygen or
coordination. fluorine). These atoms are usually difficult to locate on the powder
For these heteronuclear compounds, building an independent diffraction data because of their low atomic density (in particular
model from the homonuclear DQ build-up approach is indeed not when they are located close to a heavy atom). As will be shown
ideal as (i) the cations (Al and P) have high natural abundance (i.e., in the next step, the 17O and 19F NMR data are mostly interesting in
the isolated spin pair hypothesis is not valid), (ii) most of them are the refinement step, rather than for the structure determination
quadrupolar and therefore one would require the use of the DOR itself.
technique mentioned above. The a priori knowledge about Al–P
connectivity scheme however proved to be the useful information 2.3. Structure refinement
for structure determination, actually more important than the
distances themselves [98]. The connectivity between the Al poly- Structure refinement consists of performing small variations of
hedra and the P tetrahedra can be obtained from ‘standard’ 2D the atomic coordinates, starting from the previously determined
27
Al–31P heteronuclear NMR spectra, either J-based or dipolar- structural model, in order to provide the best match to one or
based (CP, INEPT or HMQC-type). The only requirement is, for several given cost functions (total energy or experimental data,
compounds with numerous overlapping 27Al resonances, to use a usually diffraction). For powders, the most used procedure is the
multiple-quantum (MQ) filter so as to obtain an isotropic – thus so-called Rietveld method, with powder diffraction pattern(s) as
resolved – 27Al dimension [94]. cost function.
From the 2D 27Al–31P NMR spectrum, large building units can The recent progresses in advances in calculation of the ab initio
be constructed ‘by hand’, by looking at the cross-correlation peaks, NMR parameters (full tensors of chemical shift and electric field
assuming only the first Al–P neighbors give signals, and consider- gradients, as well as their relative orientation in the unit cell) for
ing typical Al–P distances. This approach was applied to layered extended periodic systems with codes like NMR-CASTEP [98],
aluminophosphates AlPO-(Al5P7)–DAE and AlPO–(Al5P7)–DAP WIEN2K [99], ABINIT [100], Quantum-Espresso [101], etc, renders
[95,98]. When strait usage of diffraction-based algorithms was possible – in theory – the use of NMR parameters as cost function
difficult, a 100% rate of convergence was obtained when connected for a refinement (or for a structure search).
polyhedra – instead of isolated polyhedra – were used as input The proofs of such a refinement have been given on simple
objects in FOX. The choice of the initial set of structural subunits is solids, i.e., solids containing atoms sitting on positions with small
important as the more documented the knowledge about the Al–P number of variable atomic coordinates [102,103]. For example, in the
adjacency matrix is, the more efficient and fast the structural case study solid MgBr2, the position of the bromine atom along the
search will be. This is illustrated in Fig. 5 for AlPO–(Al5P7)–DAP: by c-axis of the unit cell was refined by minimizing the difference
increasing the connectivity between the identified polyhedra between experimental 81Br (nuclear spin of 3/2) quadrupolar
(which can be done by a careful analysis of the correlations on coupling constant (CQ) value and those calculated for different
the 2D NMR spectrum), the number of degrees of freedom is positions of the bromine atom along the c axis. A good agreement
reduced which in turn reduces the number of independent with the diffraction data was obtained, validating the method.
parameters to be determined from the powder diffraction data. In ITQ-4 zeolite (four T-sites), such procedure has also been carried
The immediate impacts are –(i) to bring the system below the out using both Gaussian and CASTEP codes, yielding refined struc-
threshold of the computability limit, i.e., allowing convergence to tures close to that obtained from single-crystal X-ray data [104].
be reached, and –(ii) to greatly diminish the calculation time In principle, these calculations could also be used in the
needed to converge [96,97]. It also pushes further the limits of the previous step of structure solution search, with the NMR para-
complexity of the systems that can be solved. meters as an additional cost function to the powder diffraction
pattern and total energy. Unfortunately, contrary to a theoretical
powder diffraction pattern, which takes less than a millisecond to
compute (as there is an analytical relationship between the atomic
coordinates and the powder diffraction diagram through the
structure factor), the time necessary for the calculation of the
NMR parameters for one candidate structure for inorganic systems
at the DFT-level takes minutes to hours, depending on the code
and on the complexity of the solid. Therefore, it is still impossible,
at the moment, to implement these calculations within the
diffraction softwares (direct space or charge flipping) working
with a Monte-Carlo approach, that generates billions of trial
structures to be evaluated.
Even if the NMR parameters are not used as cost function for
the refinement procedure, comparison of experimental and calcu-
lated NMR parameters can provide an assessment of the proposed
structural model. In particular, a large discrepancy very likely
indicates a problem in the position of one or more atoms, in
particular if the structure was solved from laboratory X-ray
powder diffraction data (as shown in Fig. 6 for an inorganic
fluoride). Numerous examples have shown that these discrepan-
cies can usually be corrected after a geometry optimization at the
DFT level. For details, I refer to the recent review by Bonhomme
Fig. 5. Left: Time required to get convergence to a structural model for AlPO4–
(Al5P7)–DAP by using the FOX software versus the number of degrees of freedom (n)
et al. and all the references herein [105], as well as improvement of
The various SBUs, constructed from the 27Al–31P 2D NMR correlation spectrum and the calculations reported this year for aluminophosphates and
used for the structure search are shown [88]. inorganic fluorides [106,107]. This step of structure optimization at

Please cite this article as: C. Martineau, Solid State Nucl. Magn. Reson. (2014), http://dx.doi.org/10.1016/j.ssnmr.2014.07.001i
8 C. Martineau / Solid State Nuclear Magnetic Resonance ∎ (∎∎∎∎) ∎∎∎–∎∎∎

solids, dynamics, etc. This deviation can be straightforwardly


detected when the number of NMR resonances does not match
the number of crystallographic sites in the asymmetric unit cell.
This non-periodicity has to be taken with care, in particular if the
NMR data are used in the search for the periodic model as
reported in the previous sections, as it can provide the wrong
number of inequivalent atoms.
Two types of non-periodic elements can be considered: a
variation of the atoms positions from one asymmetric unit to the
other, that generates topological differences, and a variation of the
nature of the atom occupying a given crystallographic positions
(substitution of atoms, presence of vacancies, defects), that gen-
erates chemical differences without altering the long-range topol-
ogy [111]. The latter is the most common and concern cations, e.g.,
as often encountered in zeolites or silicate (Al/Si, B/Si substitu-
tions), or the anions, e.g., as found in fluorinated metal-phosphates
or MOFs (F/OH distributions). Water molecules in porous frame-
works are usually ‘extra-framework’ species, i.e., they modify
locally the chemical composition but do not substitute a frame-
work atom, thus do not alter the overall topology of the solid. The
case of topological disorder is less usual, but was shown to occur
Fig. 6. Experimental 19F MAS NMR spectrum of Ba5Al3F19 (top) The NMR spectra in borosilicate, with a random distribution of SiO4 tetrahedra and
calculated with the NMR-CASTEP code starting with the atomic coordinates from BO3 units [13]. Information about the distribution on a long-range
(a) Rietveld refinement of the synchrotron powder diffraction data (SPD) and scale, i.e., random or correlated order, can be obtained from ss-
(b) geometry optimization (DFT) are shown below. The agreement with the
NMR data; a few recent examples are detailed below.
experimental data is better for the DFT-optimized structure.

3.1. Cation distribution


the DFT level was shown to be particularly relevant for compounds
containing atoms with very close or at the other extreme very Two types of cation distributions exist: random or locally
different electron density. In the former case (e.g., zeolites, sili- ordered, that give rise to different ss-NMR signatures. Typical
cates, etc.), the electron density map appears as a continuous examples are the Al/Si distributions in zeolites, which can be
cloud. In the latter case, the heavy atoms blow the electron density identified indirectly through the characteristic 29Si isotropic che-
map and hinder the location of the light atoms. This is for example mical shifts, which range is well-defined for SiO4 tetrahedra
the case of lead or barium fluorides [27,29,108]. It was shown that having zero to four aluminum as first cation neighbors [112]. This
an improved agreement between experimental and calculated distribution, highly important as it determines the acidic proper-
NMR parameters was obtained after geometry optimization. EFG ties of the zeolite, is known to be well ordered in a few zeolites, in
tensors and 19F isotropic chemical shifts are indeed highly sensi- particular in the natural zeolite natrolite, and it was very recently
tive to the distortion of the cationic polyhedron, and to the nature shown that the distribution could be controlled in synthetic
and number of atoms in the coordination sphere of the fluoride natrolite as well [113]. Other cations can be included in the
ions. When the calculations are performed simultaneously on framework of silicoaluminophosphate, like zinc atoms in STA-1.
several nuclei in the solid, this reinforces the validity of the In that case, the 31P nucleus, which isotropic chemical shift is sensitive
structural model. The DFT-calculations are also important for line to the number of Zn atoms surrounding the phosphorus atoms, is the
assignment, and were shown to give assignments consistent with perfect spy to identify and quantify the cation distribution [114]. It can
those independently carried out using 2D correlation experiments be extended to 29Si/11B distribution, as was illustrated for MCM-70
[109]. [115].
Such distributions also exist in dense aluminosilicate, for example
in gehlenite Ca2Al2SiO7 [116], for which Florian et al. reported a fully
3. Beyond periodicity disordered arrangement of the tetrahedral pairs in the a–b plane of
the structure. This type of information was obtained by combining a
This step concerns all parts of the crystals that cannot be battery of 1D and 2D NMR experiments, including {29Si}27Al hetero-
described by the simple periodic rules of the space groups as nuclear multiple-quantum correlation (HMQC) experiments, along
reported in the International Table of Crystallography [65]. Because with DFT-calculations of the NMR parameters.
of this lack of periodicity, powder diffraction has difficulty to address Other examples reported recently include various pyrochlores
these issues. And in fact it has long been neglected, because the ‘low’ [117,118], in particular Sn/Ti distribution in Y2Ti2  xSnxO7 [119]
resolution the NMR data were generally in good agreement with the investigated by 119Sn (I ¼1/2) NMR spectroscopy though 1D and
average structures proposed by diffraction methods. However, with 2D 119Sn–119Sn INADEQUATE measurements and DFT-calculations,
the advances in high-resolution NMR methods, structural details that or LaSrAlO4 [120], in which the La/Sr ordering was observed by
go beyond these average structures have been revealed on the NMR combined analysis of synchrotron and neutron diffraction, pair
spectra. NMR Crystallography therefore plays a crucial role in this distribution function, 27Al 1D and MQMAS NMR and atomistic
completion of the structure description, especially now that the calculations.
measurements can be associated to DFT-calculations on not fully
periodic system [110]. 3.2. Anion distribution
There are many reasons for which the structure of a compound
can deviate from strict periodicity, ranging from defect, to sub- This is the typical case of fluorine/hydroxyl substitutions. These
stitutions (which leads to chemical or topological distributions), species are isoelectronic, and therefore they cannot be distin-
stacking disorder, presence of guest molecules in porous or layers guished from X-ray diffraction data (neither single-crystal nor

Please cite this article as: C. Martineau, Solid State Nucl. Magn. Reson. (2014), http://dx.doi.org/10.1016/j.ssnmr.2014.07.001i
C. Martineau / Solid State Nuclear Magnetic Resonance ∎ (∎∎∎∎) ∎∎∎–∎∎∎ 9

Fig. 7. REDOR results and apportionments. (A) Experimental 13C{15N} REDOR curves (squares) are compared with the molecular dynamics simulations for the different linker
apportionments, large clusters (LC), small clusters (SC), random (Ran), and alternating (Alt). (B) The derived apportionments for the different MTV-MOF-5s [125].

Please cite this article as: C. Martineau, Solid State Nucl. Magn. Reson. (2014), http://dx.doi.org/10.1016/j.ssnmr.2014.07.001i
10 C. Martineau / Solid State Nuclear Magnetic Resonance ∎ (∎∎∎∎) ∎∎∎–∎∎∎

powder). Examples were reported for oxifluorides [121] and was used to identify the local geometry of the Mg and Zn atoms
fluorinated AlPOs [122,123]. Notably, the nature of the distribution [129]. In particular, the line broadenings observed on the Raman
in AlPOs (random or not) can be evaluated by recording 19F–13P and 67Zn/25Mg NMR spectra have shown an increased configura-
and/or 19F–27Al 2D NMR spectra. In ULM-3(Al), all bridging tional disorder and crystal defects in the lattice with the increase
positions between the aluminum polyhedra were found fully of the Mg substitution into a ZnO-rich lattice.
occupied by F  ions while the similar compound ULM-4(Al) a A combination of a wide range of 17O, 29Si, 25Mg, 1H and 2H
random distribution of F ions and OH groups has been clearly NMR data with first-principles calculations (CASTEP), X-ray dif-
identified. The difference in F/OH distribution was ascribed to the fraction and FTIR spectroscopy was used to identify the location of
amount of water and the pH of the synthesis medium. When all the protons and magnesium vacancies in the hydrous form of
several positions are available for fluorine, NMR can also provide wadsleyite β-(Mg,Fe)2SiO4, a mineral is found in the Earth's
information about the location. For example, in the microporous mantle. In its fully hydrated state, this mineral was thought to
AlPO4-CLO [124] solid (CLO ¼cloverite-type structure), F can either be ordered from the X-ray diffraction data. In fact, the 1H and 29Si
be trapped in double-four ring (D4R) cages forming the frame- MAS NMR spectra, supported by all the other data, showed that a
work, or be positioned at the terminal positions (Al–F, P–F) that small number of additional protons was also present as Si–OH
interrupt the framework. 2D correlation NMR spectra 19F–31P and groups located in the Mg(3) vacancies [130].
19 27
F- Al and the absence of signal in 19F–19F NMR experiments Disorder also becomes important in the case of nanoparticles;
have indicated the correlated presence between one F ion the D4R this disorder that can be analyzed by solid-state NMR. For
and only hydroxyl groups on the terminal positions of the example, in ZnSe nanoparticle [131], solid-state 77Se and 67Zn
aluminum and phosphorus polyhedra. NMR spectra reveal relationship between the distributions of 77Se
and 67Zn environments and the size of the nanoparticle, in
3.3. Linker distribution contrast with high degrees of atomic positional order established
by transmission electron microscopy and X-ray diffraction. First-
Metal-organic frameworks (MOFs) have structures based on principles calculations of NMR parameters were further used to
the assembly of cations forming clusters, linked together though distinguish between atomic positional and electronic disorder that
polycomplexing organic molecules (mostly carboxylates and azo- propagate from the nanoparticle surfaces.
lates). One key in the success of these materials is the possibility to Another possibility to probe defects in crystal is through the
tune its properties by incorporating linkers with functional groups. use of T1 relaxometry, as was shown recently for crystals of fluorite
To further increase the multi-functional character, MOFs were CaF2 [132,133]. T-1 relaxometry was shown to be sensitive to
prepared with different linkers, with the aim to obtain synergetic inhomogeneity of powders (large defect-free grains versus defec-
effects. The difficulty is to control the spatial distribution of these tive small ones). In combination with analysis of the diffraction
linkers on a mesoscale. Reimer and Yaghi have reported a 13C–15N line profile, these measurements were employed to characterize
(using selectively labeled compounds on the functional group) the lattice detect in CaF2 after extensive milling. In particular, they
REDOR study coupled to simulations of the spin system [125], to managed to distinguish the different types of defects, surface
evaluate the distribution of 1,4-benzenedicarboxylate (bdc) linkers defects (related to domain size), line defects (dislocations) and
functionalized with different groups (bdc-NH2, bdc-NO2, bdc- point (Frenkel) defects, which were found to have a quite different
(CH3)2, bdc-(OC3H5)2 and bdc-(OC7H7)2) in MOF-5. The results, effect on the powder pattern as well as on the T-1 spin-lattice
shown in Fig. 7, proves that a control of the distribution of the bdc relaxation times.
linkers, ranging from random, alternating and cluster, could be
obtained by choosing the adequate pair of linkers. 3.6. Glasses

3.4. Correlated disorder The ultimate lack of periodicity and long-range order is found
in glasses, which makes their structure determination particularly
The spatial range of slight structural disorder in partially challenging from the little information available on X-ray diffrac-
ordered solids was investigated in surfactant-templated silicate tion patterns. The combination of multinuclear solid-state NMR
layers by analysis of the lineshape of 2D DQ 29Si NMR experiments. (17O, 29Si, 11B, 27Al, 19F, etc) data with molecular dynamics (MD)
[126]. In such compounds, cross-correlations peaks with shapes simulations and calculations of NMR parameters from first-
highly elongated along the DQ diagonal of the 2D 29Si–29Si DQ-SQ principles have proven highly efficient to overcome these difficul-
INADEQUATE NMR spectrum were indeed observed. This was ties and provide accurate descriptions of the local structural
related to strong interrelationships between local framework features in oxide or halide glasses. Details about the MD-GIPAW
ordering and disordering and the mobilities of the surfactant calculations as well as numerous examples can be found in recent
headgroups. Measurement of scalar coupling distributions [127] reviews by Charpentier et al. [134–136] and references within.
can also reveal details about local disorder in solids. Another application of NMR crystallography in glasses is the
detection of polyamorphism, i.e., the simultaneous presence of
3.5. Defects different amorphous phases within a glass, which is inaccessible
from powder diffraction. The concept was illustrated in the
Defects in crystal structures are also important features for molecular glass of triphenyl phosphite with 31P NMR measure-
understanding the properties of solids, and can benefit from ments [137].
analysis of NMR spectra, as defects mostly are local-range in
nature.
For example, in the Zircon phase ScVO4  x (0.0rx r 0.1) [128], 4. Outlook
oxygen defect concentrations in bulk samples was compared with
the fully oxidized zircon structure ScVO4 using powder X-ray In this Trends papers, the various strategies that are currently
diffraction, neutron diffraction, and bulk magnetic susceptibility being developed for NMR Crystallography have been reported and
data as well as 45Sc and 51V solid-state NMR spectroscopy. In the illustrated with examples taken in the field of inorganic or hybrid
wurtzite type Zn1  xMgxO (0 rx r0.15) alloys, a combination of materials. All throughout the paper is shown the great benefit of
Raman and 67Zn/25Mg NMR spectroscopy and neutron scattering using ssNMR data in combination with powder diffraction data

Please cite this article as: C. Martineau, Solid State Nucl. Magn. Reson. (2014), http://dx.doi.org/10.1016/j.ssnmr.2014.07.001i
C. Martineau / Solid State Nuclear Magnetic Resonance ∎ (∎∎∎∎) ∎∎∎–∎∎∎ 11

and first-principles calculations to achieve a more efficient struc- [13] E. Véron, M.N. Garaga, D. Pelloquin, S. Cadars, M.R. Suchomel, E. Suard,
ture search, and to provide a more accurate structural model than D. Massiot, V. Montouillout, G. Matzen, M. Allix, Inorg. Chem. 52 (2013) 4250.
[14] K.E. Johnston, J.M. Griffin, R.I. Walton, D.M. Dawson, P. Lightfoot, S.E. Ashbrook,
if those techniques were used independently. If the determination Phys. Chem. Chem. Phys. 13 (2011) 7565.
of the unit cell dimension is still a step that is ‘diffraction specific’ [15] K.E. Johnston, C.C. Tang, J.E. Parker, K.S. Knight, P. Lightfoot, S.E. Ashbrook,
(electron or X-ray), in all the other steps, from space group J. Am. Chem. Soc. 132 (2010) 8732.
[16] D.L. Bryce, E.B. Bultz, D. Aebi, J. Am. Chem. Soc. 130 (2008) 9282.
determination, to building and refinement of the model, the [17] H. Colas, L. Bonhomme-Coury, C. Coelho Diogo, F. Tielens, F. Babonneau,
inclusion of NMR data and first-principles calculations greatly C. Gervais, D. Bazin, D. Laurencin, M.E. Smith, J.V. Hanna, M. Daudon,
increases the efficiency of the structure determination process and C. Bonhomme, CrystEngComm 15 (2013) 8840.
[18] B. Zhou, V.K. Michaelis, S. Kroeker, J.E.C. Wren, Y. Yao, B.L. Sherriff, Y. Pan,
the chance of success. Ss-NMR also allows describing crystal CrystEngComm 15 (2013) 8739.
structures beyond the rules of strict periodicity, by providing [19] S.C. Neumair, S. Vanicek, R. Kaindl, D.M. Többens, C. Martineau, F. Taulelle,
accurate details about chemical, topological disorder, dynamics, J. Senker, H. Huppertz, Eur. J. Inorg. Chem. (2011) 4147.
[20] M.D. Alba, P. Chain, T. Gonzalez-Carrascos, J. Am. Ceram. Soc. 92 (2009) 487.
etc. These techniques are also applicable to the fully disordered
[21] A.J. Fernandez-Carrion, M. Allix, P. Florian, M.R. Suchomel, A.I. Becerro,
compounds that are inorganic glasses. As mentioned in the J. Phys. Chem. C 116 (2012) 21523.
introduction, a complete description of a crystal structure requires [22] M. Allix, M.D. Alba, P. Florian, A.J. Fernandez-Carrion, M.R. Suchomel,
both these aspects to be address. This Generalized Crystallography, A. Escudero, E. Suard, A.I. Becerro, J. Appl. Crystallogr. 44 (2011) 846.
[23] H. Luhrs, A. Senyshyn, S.P. King, J.V. Hanna, H. Schneider, R.X. Fischer,
a concept introduced by MacKay twenty years ago [10] although Z. Kristallogr. 228 (2013) 457.
he was not specific about the adequate techniques to use, can be [24] R. Dervisoglu, D.S. Middlemiss, F. Blanc, L.A. Holmes, Y.L. Lee, D. Morgan,
performed from powder diffraction, ss-NMR data and DFT- C.P. Grey, Phys. Chem. Chem. Phys. 16 (2014) 2597.
[25] K.E. Johnston, M.R. Mitchell, F. Blanc, P. Lightfoot, S.E. Ashbrook, J. Phys.
calculations. Chem. C 117 (2013) 2252.
The strategies reported here are diverse in their spirit, this can [26] P.J. Pallister, I.L. Moudrakovski, G.D. Enright, J.A. Ripmeester, CrystEngComm
be explained by the large variety of existing NMR experiments. 15 (2013) 8808.
[27] C. Martineau, F. Fayon, M.R. Suchomel, M. Allix, D. Massiot, F. Taulelle, Inorg.
The general trend is nonetheless to develop specific approaches of Chem. 50 (2011) 2644.
NMR Crystallography for a given class of samples. [28] C. Martineau, C. Legein, M. Body, O. Péron, B. Boulard, F. Fayon., J. Solid State
One drawback often mentioned as a barrier for a wider use of Chem. 199 (2013) 326.
[29] C. Martineau, F. Fayon, C. Legein, J.-Y. Buzaré, M. Body, D. Massiot,
NMR is the ‘slow’ aspect of ss-NMR experiments, which, this is F. Goutenoire, Dalton Trans. (2008) 6150.
true, is much more time consuming to perform than the collection [30] M. Biswal, M. Body, C. Legein, G. Corbel, A. Sadoc, F. Boucher, J. Phys. Chem.
of a single X-ray powder diffraction pattern. However, the recent C 116 (2012) 11682.
[31] A.-L. Rollet, M. Allix, E. Véron, M. Deschamps, V. Montouillout, M.R. Suchomel,
advances in NMR signal enhancement techniques (hyperpolariza-
E. Suard, M. Barré, M. Ocana, A. Sadoc, F. Boucher, C. Bessada, D. Massiot, F. Fayon,
tion methods) as well as other methods of time reduction (non- Inorg. Chem. 51 (2012) 2272.
uniform sampling, use of parallel receiver, better detection, etc), [32] T. Krahl, M. Ahrens, G. Scholz, D. Heidemann, E. Kemnitz, Inorg. Chem. 47
have considerably reduced this time. Therefore, the comparison (2008) 663.
[33] B. Stoger, M. Weil, J. Skibsted, Dalton Trans. 42 (2013) 11672.
now is between the time necessary to perform the structure [34] S.E. Ashbrook, M. Cutajar, C.J. Pickard, R.I. Walton, S. Wimperis, Phys. Chem.
determination using powder diffraction data only (which can be Chem. Phys. 10 (2008) 5754.
infinite) and the time necessary to (i) record the NMR experiments [35] P.J. Byrne, J.E. Warren, R.E. Morris, S.E. Ashbrook, Solid State Sci. 11 (2009)
1001.
and (ii) to exploit the results for the structural analysis. The [36] M. Castro, V.R. Seymour, D. Carnevale, J.M. Griffin, S.E. Ashbrook, P.A. Wright,
balance now becomes favorable to NMR, and therefore one can D.C. Apperley, J.E. Parker, S.P. Thomspon, A. Fécan, N. Bats, J. Phys. Chem. C
expect many more of such combined contributions in the field of 114 (2010) 12698.
[37] V.R. Seymour, E. Eschenroeder, M. Castro, P.A. Wright, S.E. Ashbrook,
inorganic solids. CrystEngComm 15 (2013) 8668.
[38] M. Kanzaki, X. Xue, Inorg. Chem. 51 (2012) 6164.
[39] S.E. Ashbrook, M. Cutajar, J.M. Griffin, Z.A.D. Lethbridge, R.I. Walton,
Acknowledgment S. Wimperis, J. Phys. Chem. C 113 (2009) 10780.
[40] J. Xu, L. Chen, D. Zeng, J. Yang, M. Zhang, C. Ye, F. Deng, J. Phys. Chem. B 111
(2007) 7105.
CM is grateful for financial support from the ANR under [41] C. Serre, M. Haouas, F. Taulelle, W. Van Beek, G. Férey, C. R. Chimie 13 (2010)
contract ANR-12-JS08-0008. 336.
[42] L. Mafra, J. Rocha, C. Fernandez, G.R. Castro, S. Garcia-Granda, A. Espina,
S.A. Khainakov, J.R. Garcia, Chem. Mater. 20 (2008) 3944.
Referencesn [43] C. Martineau, T. Loiseau, L. Beitone, G. Férey, B. Bouchevreau, F. Taulelle,
Dalton Trans. 42 (2013) 422.
[44] M. Amri, S.E. Ashbrook, D.M. Dawson, J.M. Griffin, R.I. Walton, S. Wimperis,
[1] 〈http://www.iucr.org/〉. J. Phys. Chem. C 116 (2012) 15048.
[2] C. Dejoie, L.B. McCusker, C. Baerlocher, M. Kunz, N. Tamura, J. Appl. Crystal- [45] R. Gautier, R. Gautier, O. Hernandez, N. Audebrand, T. Bataille, C. Roiland,
logr. 46 (2013) 1805. E. Elkaim, L. Le Pollès, E. Furet, E. Le Fur, Dalton Trans. 42 (2013) 8124.
[3] (a) D. Harker, J.S. Kasper, Acta Cryst. 1 (1948) 70; [46] G. Tricot, O. Mentré, S. Cristol, L. Delevoye, Inorg. Chem. 51 (2012) 13108.
(b) C. Giacovazzo, Int. Tables Crystallogr. B 2.3 (2006) 210. [47] S.J. Sedlmaier, M. Doeblinger, O. Oeckler, J. Weber, J. Schmedt Auf Der Günne,
[4] (a) A.L. Patterson, Phys. Rev. 46 (1934) 372; W. Schnick, J. Am. Chem. Soc. 133 (2011) 12069.
(b) M.G. Rossmann, E. Arnold, Int. Tables Crystallogr. B 2.3 (2006) 235. [48] C. Volkringer, D. Popov, T. Loiseau, G. Férey, M. Burghammer, C. Riekel,
[5] C. Baerlocher, L.B. McCusker, L. Palatinus, Z. Kristallogr. 222 (2007) 47. M. Haouas, F. Taulelle, Chem. Mater. 21 (2009) 5695.
[6] R. Cerny, V. Favre-Nicolin, Z. Kristallogr. 222 (2007) 105. [49] C. Volkringer, T. Loiseau, M. Haouas, F. Taulelle, D. Popov, M. Burghammer,
[7] (a) M.W. Deem, J.M. Newsam, J. Am. Chem. Soc. 114 (1992) 7189; C. Riekel, C. Zlotea, F. Cuevas, M. Latroche, D. Phanon, C. Knöfelv, P.L. Llewellyn,
(b) K.D.M. Harris, M. Tremayne, P. Lightfoot, P.G. Bruce, J. Am. Chem. Soc. 116 G. Férey, Chem. Mater. 21 (2009) 5783.
(1994) 3543. [50] T. Ahnfeldt, N. Guillou, D. Gunzelmann, I. Margiolaki, T. Loiseau, G. Férey,
[8] M. Falcioni, M.W. Deem, J. Chem. Phys. 110 (1999) 1766. J. Senker, N. Stock, Angew. Chem. Int. Ed. 48 (2009) 5163.
[9] R.K. Harris, R.D. Wasylishen, M.J. Duer (Eds.), NMR Crystallography, John [51] C. Volkringer, T. Loiseau, N. Guillou, G. Férey, M. Haouas, F. Taulelle,
Wiley & Sons Ltd, Chichester, 2009. N. Audebrand, I. Margiolaki, D. Popov, M. Burghammer, C. Riekel, Cryst.
[10] (a) A.L. MacKay, Acta Crystallogr. 38 (1982) 165; Growth Des. 9 (2009) 2927.
(b) A.L. Mackay, J. Mol. Struct. 336 (1995) 293; [52] H. Reinsch, B. Marszalek, J. Wack, J. Senker, B. Gil, N. Stock, Chem. Commun.
(c) A.L. Mackay, Struct. Chem. 13 (1995) 215; 48 (2012) 9486.
(d) A.L. Mackay, Comp. Maths Appl. 12 (1986) 21. [53] H. Reinsch, M. Kruger, J. Wack, J. Senker, F. Salle, G. Maurin, N. Stock,
[11] J.A. Ripmeester, R.E. Wasylishen, CrystEngComm 15 (2013) 8598. Microporous Mesoporous Mater. 157 (2012) 50.
[12] C. Martineeau, J. Senker, F. Taulelle, Ann. Rep. NMR Spectrosc. 82 (2014) 1. [54] C. Volkringer, T. Loiseau, G. Férey, C.M. Morais, F. Taulelle, V. Montouillout,
D. Massiot, Microporous Mesoporous Mater. 105 (2007) 111.
[55] J.P.S. Mowat, S.R. Miller, A.M.Z. Slawin, V.R. Seymour, S.E. Ashbrook,
n
Articles of general interest. P.A. Wright, Microporous Mesoporous Mater. 142 (2011) 322.

Please cite this article as: C. Martineau, Solid State Nucl. Magn. Reson. (2014), http://dx.doi.org/10.1016/j.ssnmr.2014.07.001i
12 C. Martineau / Solid State Nuclear Magnetic Resonance ∎ (∎∎∎∎) ∎∎∎–∎∎∎

[56] M. Mazaj, G. Mali, M. Rangus, E. Zunkovic, V. Kaucic, N.Z. Logar, J. Phys. Chem. [103*] C.M. Widdifield, D.L. Bryce, Phys. Chem. Chem. Phys. 11 (2009) 7120 (a paper
C 117 (2013) 7552. illustrating the principles of structure refinement from first-principle
[57] K.M.N. Burgess, Y. Xu, M.C. Leclerc, D.L. Bryce, Inorg. Chem. 53 (2014) 552. calcualtions of NMR parameters on MgBr2).
[58] C. Martineau, A. Cadiau, B. Bouchevreau, J. Senker, F. Taulelle, K. Adil, Dalton [104] D.H. Brouwer, I.L. Moudrakowski, R.J. Darton, R.E. Morris, Magn. Reson.
Trans. 41 (2012) 6232. Chem. 48 (2010) S113.
[59] S. Devautour-Vinot, G. Maurin, C. Serre, P. Horcajada, D.P. Da Cunha, [105*] C. Bonhomme, C. Gervais, F. Babonneau, C. Coelho, F. Pourpoint, T. Azais, S.
V. Guillerm, E. de Souza Costa, F. Taulelle, C. Martineau, Chem. Mater. 24 E. Ashbrook, J.M. Griffin, J.R. Yates, F. Mauri, C.J. Pickard, Chem. Rev. 112
(2012) 2168. (2012) 5733 (a review on the use of first-principle calculations applied to a
[60] Q. Yang, S. Vaesen, F. Ragon, A.D. Wiersum, D. Wu, A. Lago, T. Devic, wide variety of materials).
C. Martineau, F. Taulelle, P.L. Llewellyn, C. Serre, H. Jobic, C. Zhong, G. De [106] S. Sneddon, D.M. Dawson, C.J. Pickard, S.E. Ashbrook, Phys. Chem. Chem.
Weireld, G. Maurin, Angew. Chem. Int. Ed. 52 (2013) 10316. Phys. 16 (2014) 2660.
[61] P. He, J. Xu, V.V. Terskikh, A. Sutrisno, H.-Y. Nie, Y. Huang, J. Phys. Chem. C 117 [107] A. Sadoc, M. Biswal, M. Body, C. Legein, F. Boucher, D. Massiot, F. Fayon, Solid
(2013) 16853. State Nucl. Magn. Reson. 59 (2014) 1.
[62] Structure Determination from Powder Diffraction Data, W.I.F. David et al., ed. [108] M. Body, C. Legein, J.-Y. Buzaré, G. Silly, P. Blaha, C. Martineau, F. Calvayrac,
IUCr Monographs on Crystallography 13, Oxford University Press, 2002. J. Phys. Chem. A 111 (2007) 11873.
[63] International Tables for Crystallography Vol B. [109] C. Martineau, C. Legein, J.-Y. Buzaré, F. Fayon, Phys. Chem. Chem. Phys. 11
[64] F. Taulelle, Solid State Sci. 6 (2004) 1053. (2009) 950.
[65] F.W. Karau, L. Seyfarth, O. Oeckler, J. Senker, K. Landskron, W. Schnick, Chem. [110*] S.E. Ashbrook, D.M. Dawson, Acc. Chem. Res. 46 (2013) 1964 (a paper on the
Eur. J. 13 (2007) 6841. use of FT calculations for non-periodic systems).
[66] K.M.N. Burgess, I. Korobkov, D.L. Bryce, Chem. Eur. J. 18 (2012) 5748. [111*] D. Massiot, R.J. Messinger, S. Cadars, M. Deschamps, V. Montouillout,
[67] K.E. Johnston, J.M. Griffin, R.I. Walton, D.M. Dawson, P. Lightfoot, S.E. Ashbrook., N. Pellerin, E. Véron, M. Allix, P. Florian., F. Fayon, Acc. Chem. Res. 46
Phys. Chem. Chem. Phys. 13 (2011) 7565. (2013) 1975 (a nicely illustrated discussion on topological and chemical
[68] M.B. Park, A. Vicente, C. Fernandez, S.B. Hong, Phys. Chem. Chem. Phys. 15 disorder in crystalline solids).
(2013) 7604. [112] G. Engelhardt, D. Michel, High-Resolution Solid-State NMR of Silicates and
[69] B.E.G. Lucier, K.E. Johnston, W. Xu, J.C. Hanson, S.D. Senanayake, S. Yao, M. Zeolites, John Wiley & Sons, New York, 1987.
W. Bourassa, M. Srebro, J. Autschbach, R.W. Schurko, J. Am. Chem. Soc. 136 [113] J. Shin, N.H. Ahn, M.A. Camblor, C.M. Zicovich-Wilson, S.B. Hong, Chem. Mater
(2014) 1333. 26 (2014) 3361.
[70] B. Elena, G. Pintacuda, N. Mifsud, L. Emsley, J. Am. Chem. Soc. 128 (2006) [114] A.B. Pinar, L. Gomez-Hortiguela, L.B. McCusker, J. Perez-Pariente, Dalton
9555. Trans. 40 (2011) 8125.
[71] F.A. Perras, I. Korobkov, D.L. Bryce, CrystEngComm 15 (2013) 8727. [115] D. Xie, L.B. McCusker, C. Baerlocher, L. Gibson, A.W. Burton, S.-J. Hwang,
[72] D.H. Brouwer, G.D. Enright, J. Am. Chem. Soc. 130 (2008) 3095. J. Phys. Chem. C 113 (2009) 9845.
[73] D.H. Brouwer, I.L. Moudrakovski, R.J. Darton, R.E. Morris, Magn. Reson. Chem. [116] P. Florian, E. Véron, T.F.G. Green, J.R. Yates, D. Massiot, Chem. Mater. 24 (2012)
48 (2010) S113. 4068.
[74] S. Cadars, D.H. Brouwer, B.F. Chmelka, Phys. Chem. Chem. Phys. 11 (2009) [117] K.R. Whittle, G.R. Lumpkin, S.E. Ashbrook, J. Solid State Chem. 179 (2006) 512.
1825. [118] M.H. Harunsani, D.I. Woodward, M.D. Peel, S.E. Ashbrook, R.I. Walton, J. Solid
[75] Y.-C. Pan, H.-Y. Wu, G.-L. Jheng, H.-H.G. Tsai, H.-M. Kao, J. Phys. Chem. C 113 State Chem. 207 (2013) 117.
(2009) 2690. [119] M.R. Mitchell, S.W. Reader, K.E. Johnston, C.J. Pickard, K.R. Whittle, S.E. Ashbrook,
[76] D.P. Serrano, J. Aguado, G. Morales, J.M. Rodriguez, A. Peral, M. Thommes, Phys. Chem. Chem. Phys. 13 (2011) 488.
J.D. Epping, B.F. Chmelka, Chem. Mater. 21 (2009) 641. [120] C. Tealdi, C. Ferrara, L. Malavasi, P. Mustarelli, C. Ritter, A. Spinella, D. Massiot,
[77] M.W. Müller, D. Hirsemann, F. Haarmann, J. Senker, J. Breu, Chem. Mater. 22 P. Florian, J. Mater. Chem. 22 (2012) 10488.
(2010) 186. [121] J.M. Griffin, J.R. Yates, A.J. Berry, S. Wimperis, S.E. Ashbrook, J. Am. Chem. Soc.
[78] R.M. Shayib, N.C. George, R. Seshadri, A.W. Burton, S.I. Zones, B.F. Chmelka, 132 (2010) 15651.
J. Am. Chem. Soc. 133 (2011) 18728. [122] C. Martineau, B. Bouchevreau, R. Siegel, J. Senker, A. Ristic, F. Taulelle, J. Phys.
[79] F. Gramm, C. Baerlocher, L.B. McCusker, S.J. Warrender, P.A. Wright, B. Han, S. Chem. C 116 (2012) 21489.
B. Hong, Z. Liu, T. Oshuna, O. Terasaki, Nature 444 (2006) 79. [123] C. Martineau, C. Mellot-Draznieks, F. Taulelle, Phys. Chem. Chem. Phys. 13
[80] C. Baerlocher, L.B. McCusker, L. Palatinus, Science 315 (2007) 1113. (2011) 18078.
[81] C. Baerlocher, D. Xie, L.B. McCusker, S.-J. Hwang, I.Y. Chan, K. Ong, A. [124] C. Martineau, B. Bouchevreau, Z. Tian, S.-J. Lohmeier, P. Behrens, F. Taulelle,
W. Burton, S.I. Zones, Nat. Mater 7 (2008) 631. Chem. Mater. 23 (2011) 4799.
[82] D.H. Brouwer, J. Magn. Reson. 194 (2008) 136. [125] X. Kong, H. Deng, F. Yan, J. Kim, J.A. Swisher, B. Smit, O.M. Yaghi, J.A. Reimer,
[83] D.H. Brouwer, J. Am. Chem. Soc. 130 (2008) 6306. Science 341 (2013) 882.
[84] D.H. Brouwer, S. Cadars, J. Eckert, Z. Liu, O. Terasaki, B.F. Chmelka, J. Am. [126] S. Cadars, N. Mifsud, A. Lesage, J.D. Epping, N. Hedin, B.F. Chmelka, L. Emsley,
Chem. Soc. 135 (2013) 5641. J. Phys. Chem. C 112 (2008) 9145.
[85] D.H. Brouwer, Solid State Nucl. Magn. Reson. 51 (2013) 37. [127] S. Cadars, A. Lesage, M. Trierweiler, L. Heux, L. Emsley, Phys. Chem. Chem.
[86] V.A. Blatov, IUCr Comput. Commis. Newslett. 7 (2006) 4 〈http://www.topos. Phys. 9 (2006) 92.
ssu.samara.ru〉. [128] S.P. Shafi, M.W. Kotyk, L.M.D. Cranswick, V.K. Michaelis, S. Kroeker,
[87] B. Bouchevreau, Ph.D. thesis, University of Versailles, France, 2013. M. Bieringer, Inorg. Chem. 48 (2009) 10553.
[88*] D.H. Brouwer, K.P. Langendoen, CrystEngComm 15 (2013) 748 (a pedagogical [129] Y.-I. Kim, S. Cadars, R. Shayib, T. Proffen, C.S. Feigerle, B.F. Chmelka,
paper detailing the relationship between graph theory and NMR, and R. Seshadri, Phys. Rev. B 78 (2008) 195205.
[130] J.M. Griffin, A.J. Berry, D.J. Frost, S. wimperis, S.E. Ashbrook, Chem. Sci. 4
illustrated with zeolite frameworks).
(2013) 1523.
[89] A. Le Bail, Mater. Sci. Forum 378 (2001) 65.
[131] S. Cadars, B.J. Smith, J.D. Epping, S. Acharya, N. Belman, Y. Golan, B.F. Chmelka,
[90] V. Favre-Nicolin, R. Cerny, J. Appl. Crystallogr. 35 (2002) 734.
Phys. Rev. Lett. 103 (2009) 136802.
[91] S. Pagola, P.W. Stephens, J. Appl. Crystallogr. 43 (2010) 370.
[132] M. Abdellatief, M. Abele, M. Leoni, P. Scardi, J. Appl. Crystallogr. 46 (2013)
[92] R.W. Grosse-Kunstleve, et al., J. Appl. Crystallogr 30 (1997) 985.
1049.
[93] S. Sene, B. Bouchevreau, C. Martineau, C. Gervais, C. Bonhomme, P. Gaveau,
[133] M. Abdellatief, M. Abele, M. Leoni, P. Scardi, Thin Solid Films 530 (2013) 44.
F. Mauri, S. Bégu, P. Hubert Mutin, M.E. Smith, D. Laurencin, CrystEngComm
[134] A. Soleilhavoup, J.M. Delaye, F. Angeli, D. Caurant, T. Charpentier, Magn.
15 (2013) 8763.
Reson. Chem. 48 (2010) S159.
[94] C. Martineau, B. Bouchevreau, F. Taulelle, J. Trébosc, O. Lafon, J.P. Amoureux,
[135] T. Charpentier, Solid State Nucl. Magn. Reson. 40 (2011) 1.
Phys. Chem. Chem. Phys. 14 (2012) 7112.
[136] T. Charpentier, M.C. Menziani, A. Pedone, RSC Adv. 3 (2013) 10550.
[95] B. Bouchevreau, C. Martineau, C. Mellot-Draznieks, A. Tuel, M.R. Suchomel,
[137] J. Senker, J. Sehnert, S. Correll, J. Am. Chem. Soc. 127 (2005) 337.
J. Trébosc, O. Lafon, J.P. Amoureux, F. Taulelle, Chem. Mater. 25 (2013) 2227.
[96*] B. Bouchevreau, C. Martineau, C. Mellot-Draznieks, A. Tuel, M.R. Suchomel,
J. Trébosc, O. Lafon, J.P. Amoureux, F. Taulelle, Chem. Eur. J. 19 (2013) 5009
(a paper showing how to use the NMR data to drive the real space methods
on powder diffraction patterns).
[97] F. Taulelle, B. Bouchevreau, C. Martineau, CrystEngComm 15 (2013) 8613. Charlotte Martineau, b 1983, received her Ph.D. in Chemistry at the Université du
[98] M.D. Segall, P.L.D. Lindan, M.J. Probert, C.J. Pickard, P.J. Hasnip, S.J. Clark, Maine (France) in 2008. After one year as a postdoctoral researcher at the
M.C. Payne., J. Phys.: Condens. Matter 14 (2002) 2717. University of Manitoba (Canada), she became Maitre de Conférence at the
[99] P. Blaha, K. Schwarz, G.K.H. Masden, D. Kvasnicka, J. Luitz, WIEN2k (2001). University of Versailles (France) in 2009, and obtained her Habilitation à Diriger
[100] X. Gonze, J.-M. Beuken, R. Caracas, F. Detraux, M. Fuchs, G.-M. Rignanese, les Recherches (HDR) in 2013. Her research focuses on the development and
L. Sindic, M. Verstraete, G. Zerah, F. Jollet, M. Torrent, A. Roy, M. Mikami, application of methods dedicated to solving crystalline structures from solid-state
P.h. Ghosez, J.-Y. Raty, D.C. Allan, Comput. Mater. Sci. 25 (2002) 478. NMR and powder diffraction data. A particular interest is given to the study of
[101] P. Giannozzi, et al., J. Phys.: Cond. Matter 21 (2009) 395502. the non-periodic parts of hybrid and fluorinated materials, accessible by NMR
[102] F.A. Perras, D.L. Bryce, J. Phys. Chem. C 116 (2012) 19472. Crystallography.

Please cite this article as: C. Martineau, Solid State Nucl. Magn. Reson. (2014), http://dx.doi.org/10.1016/j.ssnmr.2014.07.001i

Você também pode gostar