Startseite Neutron diffraction: a primer
Artikel Open Access

Neutron diffraction: a primer

  • Richard Dronskowski EMAIL logo , Thomas Brückel , Holger Kohlmann , Maxim Avdeev , Andreas Houben , Martin Meven , Michael Hofmann , Takashi Kamiyama , Mirijam Zobel , Werner Schweika , Raphaël P. Hermann und Asami Sano-Furukawa
Veröffentlicht/Copyright: 29. April 2024

Abstract

Because of the neutron’s special properties, neutron diffraction may be considered one of the most powerful techniques for structure determination of crystalline and related matter. Neutrons can be released from nuclear fission, from spallation processes, and also from low-energy nuclear reactions, and they can then be used in powder, time-of-flight, texture, single crystal, and other techniques, all of which are perfectly suited to clarify crystal and magnetic structures. With high neutron flux and sufficient brilliance, neutron diffraction also excels for diffuse scattering, for in situ and operando studies as well as for high-pressure experiments of today’s materials. For these, the wave-like neutron’s infinite advantage (isotope specific, magnetic) is crucial to answering important scientific questions, for example, on the structure and dynamics of light atoms in energy conversion and storage materials, magnetic matter, or protein structures. In this primer, we summarize the current state of neutron diffraction (and how it came to be), but also look at recent advances and new ideas, e.g., the design of new instruments, and what follows from that.

1 Introduction

From a higher perspective, it may be considered a happy coincidence or a simple scientific necessity to combine the sexiest physical probe of all times (the neutron, neutral but with a mass and a spin) with the most powerful structure determination method for crystalline systems (diffraction), so the special importance of neutron diffraction does not have to be explained at all. The neutron, discovered almost a century ago, came at the right moment, so to speak, and in the course of time the importance of the method grew continuously. Originally a by-product of nuclear fission, neutrons are now produced selectively, on purpose, in a variety of ways, and there are entirely new processes for neutron production on the horizon, designed for the crucial experiment.

Traditionally, crystalline powders have been studied with neutron beams, but today high-precision structure determinations of single crystals are just as common as studies targeting preferential orientation (going under the term texture). Time-of-flight neutron data allow further insights based on higher resolution and improved data structures. In this respect, the application of neutron diffraction serves as a magnificent example of the convergence of science. Consistently, neutron diffraction penetrates into additional domains, thus looking at diffuse scattering needed to investigate the structural behavior of, say, nanocrystalline or not-so-crystalline matter. Likewise, neutron diffraction helps to elucidate various magnetic phenomena, in particular with instruments utilizing the neutron spin polarization, a true core competence among the structure-elucidating (and structure-imaging) methods. When one can rely on sufficient intensities and brilliance at extremely powerful sources, real-time investigation of samples is also within reach, not only for parametric studies of thermodynamic but also chemical variables. In particular, modern high-pressure research is almost unthinkable without neutron diffraction when dealing with light atoms or those looking too similar using X-ray eyes.

The prospects are excellent for neutron diffraction. Please join us for a scientific stroll through this field, written for curious beginners, but hopefully of interest for long-time practitioners as well.

1.1 Provision of neutron beams

Neutrons are abundant in nature because, together with protons, they make up atomic nuclei. With the exception of hydrogen, light atoms contain an approximately equal number of protons and neutrons, while the heavier atoms have more neutrons than protons. That being said, it does not surprise that about 40 % of the mass of your body is made up of neutrons, tightly stored in the nuclei. Neutron diffraction requires large machines, however, simply because of the strong force of particle physics which must be overcome to release neutrons from the nuclei. There are three major fundamental processes that underlie neutron scattering facilities, as depicted in Figure 1: fission, spallation, and low-energy nuclear reactions induced by charged particles. 1

Figure 1: 
Illustration of the three major fundamental processes providing neutrons for scattering instruments: fission (top), spallation (middle) and low energy nuclear reactions (bottom). Adapted from ref 2 by kind permission of Wiley-VCH.
Figure 1:

Illustration of the three major fundamental processes providing neutrons for scattering instruments: fission (top), spallation (middle) and low energy nuclear reactions (bottom). Adapted from ref 2 by kind permission of Wiley-VCH.

To begin with, fission takes place as a chain reaction in research reactors when a fissile nucleus, usually 235U, captures a neutron of the right speed. The nucleus then splits into two fragments of almost equal mass, releasing on average about 2.5 neutrons in the process. One neutron is needed to sustain the chain reaction after it has been slowed down (or moderated, another way of saying it) because fission neutrons have a large average energy of 2 MeV.

Second, there are accelerator-driven spallation neutron sources in which heavy metal targets (typically U, W, Ta, Pb, Hg) are bombarded with protons at huge energies around or above 1 GeV. After collision, the highly excited nuclei emit neutrons, protons and also pions, which collide again to produce more excited nuclei from which particles (mainly neutrons) are emitted. Typically, 10 to 30 neutrons per proton are released in the entire process. Most of these neutrons are so-called “evaporation neutrons” with average energies of about 2 MeV but there is also a fraction of high-energy cascade neutrons with GeV energies up to the incident proton energy. While spallation is the most efficient process for releasing neutrons, much heavier shielding is required due to the high-energy particles and, secondary, the high peak flux.

Third, nuclear reactions at accelerator-driven neutron sources at energies below 100 MeV have rather low neutron yields of about 10−2 to a few 10−1 neutrons per collision. At energies below 30 MeV – the domain of so-called Compact Accelerator-driven Neutron Sources (CANS) – light metal targets (Li, Be) are most suitable whereas at higher energies, proton reactions in heavy elements are more efficient. 3 Thanks to recent technological developments, such High-Current Accelerator-driven Neutron Sources (HiCANS) can provide performance on par with fission or spallation sources. 4 Accelerator-driven sources have a decisive advantage to reactor sources because neutrons can be produced on demand. With the time structure of a pulsed source, and the relation between wavelength and time-of-flight of the neutrons, one can use the entire spectrum instead of a monochromatic beam. It is then the peak flux rather than only the time average flux which matters. More accurately, rather than flux we may consider the (peak) brilliance , with a unit of n s−1 cm−2 sr−1 Å−1, quantifying how many neutrons are created per time interval and wavelength band and passing through an area and solid angle element, which ultimately determines the useful neutrons that can be scattered by the sample.

With neutron diffraction we want to measure the distances between atoms, which requires a “ruler” of similar size. This means that the neutron wavelength should be comparable to interatomic distances of about 0.1 nm (=10−10 m = 1 Å, the so-called Ångström crystallographic unit which comes in handy). With the symbol λ for the neutron wavelength, p for the neutron momentum, m for the neutron mass, v for the neutron velocity and h for the Planck constant, the neutron kinetic energy E = p 2 2 m can be related to the neutron wavelength using the De Broglie relation λ = h p = h m v . In practical units this reads

(1) λ [ nm ] 0.9045 E [ meV ]  or  E [ meV ] 0.8181 / λ [ nm ] 2

Thus, neutrons of wavelength 1 Å have a kinetic energy of about 82 meV. This fits well with typical excitation energies in solids, making neutrons also a perfect probe for studying elementary excitations in solids.

But now there is a problem: all of the above processes release neutrons with energies in the MeV range, while for neutron diffraction we need neutrons with kinetic energies about 7–8 orders of magnitude smaller. That is to say that the neutrons from the primary reactions have to be slowed down to these energies, which is usually achieved by collisions with light elements, mainly hydrogen, in so-called moderators, as already said before. The thermal equilibrium between the neutrons and the nuclei of the moderator material produces a Maxwellian spectrum with an average neutron energy determined by the temperature of the moderator. Depending on the temperature of the moderator, one speaks of “hot”, “thermal” and “cold” neutrons, as summarized in Table 1 and illustrated in Figure 2. These terms refer to the temperature equivalent of the neutrons’ kinetic energy according to the Boltzmann conversion E kin = k BT; note that 1 meV corresponds to about 11.6 K. The moderator is surrounded by a reflector, e.g., beryllium or lead, which effectively scatters the escaping higher energy neutrons back into the moderator for further thermalization.

Table 1:

Average neutron energies and wavelengths for different moderators.

Moderator Example Temperature (K) Neutron energies (meV) Wavelengths (nm)
Hot source Graphite C 2300 100–1000 0.03–0.09
Thermal source Water H2O 300 5–100 0.09–0.4
Cold source Liquid D2 25 0.05–5 0.4–4
Figure 2: 
Maxwellian distribution of neutrons for several temperatures of the medium, which the beam from the reactor core passes through (moderator); grey shaded area is an example of a neutron wavelength band selected by a monochromator.
Figure 2:

Maxwellian distribution of neutrons for several temperatures of the medium, which the beam from the reactor core passes through (moderator); grey shaded area is an example of a neutron wavelength band selected by a monochromator.

Both the extraction and transport of neutrons from the moderators is done through evacuated beam tubes, which allow the neutrons to leave the moderator and pass through the biological shielding to the neutron instrument.

1.2 Neutron–matter interaction

Somewhat simplified, X-rays, electrons, and neutrons are the main probes for studying matter at atomic length scales. Because X-rays and electrons interact with matter mainly through the strong electrostatic Coulomb interaction, this leads to a limited penetration depth simply due to scattering from the electron cloud of the atoms. For neutrons, being neutral particles with no electrostatic Coulomb interaction, however, condensed matter appears quite empty, as the main interaction with atoms is mediated through the short-range strong force with atomic nuclei, and they are about five orders of magnitude smaller than the size of an atom’s electron cloud (= the size of the atom). This explains why most materials are fairly transparent to neutrons.

The strong interaction between neutrons and nuclei is an extremely complex particle physics problem. Nonetheless, Nobel laureate Enrico Fermi took advantage of the point-like nature of nuclei being so much smaller than the wavelength of thermal neutrons to describe the interaction, and this led to the so-called Fermi pseudopotential for nuclear scattering: 5

(2) V ( r ) = 2 π 2 m b δ ( r )

Here, r is the distance between the neutron and the nucleus, m is the mass of the neutron, and δ is the delta-function reflecting the effectively vanishing interaction range. The pre-factor in (2) is chosen such that the total cross section σ tot, i.e., the total number of neutrons scattered per second normalized to the flux of incident neutrons, is given by:

(3) σ tot = 4 π b 2

Since a point-like scatterer has no internal structure, the scattering from it is isotropic, i.e., nuclear scattering has no angular dependence. The important quantity in (2) and (3) is the bound scattering length b, which is generally a complex number in need of explanation. While the real part is a measure of the strength of the interaction, the imaginary part measures absorption. Moreover, b is a phenomenological parameter that varies rather randomly throughout the periodic table, so it must be determined experimentally, for values see ref 6. The real part of b is positive for a repulsive pseudopotential, the normal case. In rare but important cases it can be negative, however, corresponding to an attractive potential. In addition, b is isotope specific and depends on nuclear spin states: b + when the neutron (with spin = ½) and nuclear spin (I) are aligned (total spin = I+½), b for total spin = I−½.

The fact that the distribution of isotopes and nuclear spin orientations is usually random has the important consequence that scattering can be separated into a coherent and incoherent part. For coherent scattering, the scattered waves from different scattering centers interfere, i.e., a regular superposition of the relative phases occurs. In accordance with (3), coherent scattering depends on the average value b coh = 〈b〉 of the scattering length of the nuclei:

(5) σ coh = 4 π b 2

In a crystal with lattice periodicity (that is, a regular crystal with crystallographic symmetry), the coherent scattering length is responsible for the Bragg scattering, just like for the typical X-ray case. Incoherent scattering generates noise, however, because it arises from the random distribution of the deviations of the scattering length from their mean value and does not show interference effects:

(6) σ inc = 4 π { b 2 b 2 }

An important and real-world example is given by elemental hydrogen 1 1H, which has a nuclear spin of ½. With b + = 10.82 fm and b = −47.42 fm one obtains b coh = 1 4 b + 3 4 b + = 3.74  fm and σ coh = 1.76 barn, σ inc = 4 π ( 1 4 b 2 + 3 4 b + 2 ) = 80.27  barn (1 barn = 100 fm2 = 10−28 m2 = 10−24 cm2). So, the most abundant hydrogen 1 1H has a small coherent but a huge incoherent cross section, and this leads to a high flat background for scattering from samples containing hydrogen. Fortunately enough, deuterium 2 1H ≡ D, on the other hand, has cross sections of σ coh = 5.59 barn, σ inc = 2.05 barn. Hence, the much smaller incoherent scattering cross section of D can be used to minimize background otherwise present due to hydrogen’s incoherent scattering and allows for contrast variation by isotope substitution.

Figure 3 displays neutron cross sections for selected elements. While the cross section for high-energy X-rays is (to first approximation) proportional to the number of electrons (or the atomic number Z) squared and varies continuously throughout the periodic table, this is very different for neutrons. Hydrogen is virtually invisible for X-rays in the presence of heavier elements, and neighboring elements like Mn, Fe or Ni are hardly distinguishable for high-energy X-rays. Not so for neutron diffraction, thereby offering a clear advantage of neutrons.

Figure 3: 
Coherent scattering cross sections of selected elements throughout the periodic table for neutrons. The area of the circle is proportional to the cross section. The colors blue and green indicate the sign of the coherent scattering length (positive and negative, respectively). For comparison, the X-ray Thomson scattering cross section for a single electron is 0.665 barn. The classical coherent X-ray cross section at high energies varies smoothly with the square of the number of electrons throughout the periodic table.
Figure 3:

Coherent scattering cross sections of selected elements throughout the periodic table for neutrons. The area of the circle is proportional to the cross section. The colors blue and green indicate the sign of the coherent scattering length (positive and negative, respectively). For comparison, the X-ray Thomson scattering cross section for a single electron is 0.665 barn. The classical coherent X-ray cross section at high energies varies smoothly with the square of the number of electrons throughout the periodic table.

For magnetic materials, an additional scattering mechanism besides nuclear scattering must be considered. Magnetic scattering is due to the electromagnetic interaction with the internal magnetic field, resulting from the magnetic moment associated with the neutron’s nuclear spin. Fortunately, magnetic scattering is comparable in magnitude to nuclear scattering: the magnetic scattering length of neutrons from a single unpaired electron amounts to 1 2 γ n r 0 2.67 fm, where γ n denotes the magnetic dipole moment of the neutron expressed in nuclear magnetons and r 0 is the classical electron radius. Since magnetism is a vector quantity, magnetic neutron scattering is strongly dependent on the direction of the magnetic moment and the polarization of the neutron beam. 5 Also, because scattering occurs from unpaired electrons of the atom they belong to, magnetic neutron scattering exhibits a form factor dependence similar to the form factor of X-rays. Unlike nuclear scattering, which is independent of the scattering angle (see above), the intensity of magnetic scattering from an atom decreases rapidly with increasing scattering angle. In fact, it decreases even faster than for X-rays because the unpaired electrons are in the atom’s outer region, so magnetic Bragg peaks are found for small diffraction angles, in the “early” region of the standard diffraction pattern. Magnetic form factors are tabulated in refs 6 or 7.

Let’s turn to the imaginary part of the scattering length, connected with neutron absorption. Most elements have a very small neutron absorption cross section which has been tabulated, 6 proportional to the neutron wavelength:

(7) σ a λ 1 v

It can be argued that the wavelength λ and the neutron velocity v are inversely proportional, so that for longer wavelengths, i.e., lower velocities, the neutron stays near the nucleus for a correspondingly longer time, leading to a higher absorption cross section. There are some elements, however, that exhibit strong neutron absorption due to nuclear resonance processes which lead to nuclear reactions releasing charged particles, for example in 3He, 6Li, and 10B. Therefore, these reactions form the basis of neutron detectors. Strong resonance absorption with emission of gamma rays occurs, for example, in the elements Cd and Gd, which are often used in neutron shielding materials. However, care must be taken because such resonances are strongly energy dependent. For example, while Cd has an absorption cross section of about 20 kbarn at a wavelength of 0.064 nm (0.64 Å), this drops to only 8 barn at 0.02 nm (0.2 Å). In practice, absorption may even spoil a neutron diffraction experiment, which can be circumvented by combined use of annular samples, optimized incident neutron wavelength and high-intensity diffractometers or, alternatively, by using less absorbing isotopes, 8 or flat-plate sample holders. 9

2 Diffraction techniques

For neutron powder diffraction, there is the choice of instruments and techniques, each with their specific advantages, either using a monochromatic, angle-dispersive scheme, similar to the common X-ray laboratory sources, or using a pulsed, polychromatic scheme, which opens another dimension of information, as schematically depicted in Figure 4. The crystalline nature shows up by a diffraction “fingerprint” of whatever material with different degrees of long-range ordering, e.g., crystallinity, for which each line represents the Bragg law being fulfilled given some lattice spacing dubbed as d. In the subsequent sections, our course in this primer will start with monochromatic, that is, constant wavelength (CW) instruments at reactor sources, then followed by the development of time-of-flight (TOF) techniques at pulsed neutron sources. As depicted in Figure 4, a pulsed wavelength spectrum as used in TOF diffraction allows to measure the entire pattern by a single detector at a fixed angle (vertical gray lines), even with the smallest gauge volume, i.e., the smallest probed sample volume, under 90° scattering angle. For CW, a range of reflections is probed using detectors with larger angular coverage or by moving the detector with 2θ (horizontal gray line). A large range of higher diffraction angles are not of interest due to inherently low resolution, because the slopes at the intersection of the horizontal line (probe) with the reflections following a sine function scaled by d according to Bragg’s law are very similar. In contrast, TOF particularly excels at high d resolution naturally achieved in the backscattering range (high 2θ), where the vertical lines (probe) intersect almost orthogonally with the sine functions. 10

Figure 4: 
The many varieties of neutron diffraction as viewed for the exemplary case of crystalline silicon. The Bragg powder lines for constant lattice spacing d being probed by classical constant wavelength (horizontal) and time-of-flight diffraction (vertical) as well as diffraction-focusing approaches probing both dimensions; see also text.
Figure 4:

The many varieties of neutron diffraction as viewed for the exemplary case of crystalline silicon. The Bragg powder lines for constant lattice spacing d being probed by classical constant wavelength (horizontal) and time-of-flight diffraction (vertical) as well as diffraction-focusing approaches probing both dimensions; see also text.

The analysis of powder diffraction data for samples with unknown structure may be roughly split into three steps. First, characteristic d reflections need to be identified for a new phase. In the so-called indexing procedure, indices (hkl) describing sets of equivalent lattice planes in reciprocal space need to be assigned to each reflection. By carefully checking extinction rules, possible Laue class predictions with corresponding lattice parameters are derived, eased by various algorithms and computer programs. In the second step, atomic positions in the desired crystal system need to be proposed (the so-called phase problem), e.g., based on chemical intuition, other supplementary information, or by applying real space or reciprocal space methods, e.g., simulated annealing, genetic algorithms, direct methods, etc. Once one or more reasonable structural proposals are available, the traditional strength of powder diffraction, i.e., the Rietveld profile refinement method, allows to elaborate the best structural description. 11 By least-squares fitting an entire simulated diffraction pattern composed of different contributions (Bragg reflections, background, sample effects, instrumental effects, secondary phases) can be numerically compared against measured data, thereby touching upon even the most subtle structural differences. Thus, it is important to carefully select the type of analytical profile function and its parameterization in order to achieve an optimum match between the simulated and the measured profile and, hence, extract proper integrated intensities per each (also overlapping) reflection. As numerous as the instrument and sample effects affecting the profile shapes are, there are also several common profile functions. One is the ubiquitous pseudo-Voigt (pV) function, a linear combination of a Gaussian and Lorentzian function, which allows much faster computation compared to the convoluted analog, i.e., the Voigt function. Many others are even more sophisticated, e.g., the Thompson–Cox–Hastings pV dealing with axial divergence or pV including a couple of back-to-back exponentials (btb) or the Ikeda–Carpenter function used to better describe the source characteristics at spallation-source TOF diffractometers.

Traditionally, the standard powder analysis proceeds by a Rietveld profile refinement of one-dimensional data, hence TOF data are often grouped by detector banks with similar resolution. Nonetheless, more advanced methods are progressing and being optimized to exploit two-dimensional data sets as well. And yet, there is a third observable dimension completing the access to reciprocal space. Position-sensitive area detectors may collect efficiently not only more neutrons but reveal possible preferred powder orientations, texture properties and, finally, individual Bragg peaks of single crystals. While the next three chapters will concentrate on powder diffraction in more detail, chapters about texture, single crystal diffraction and further techniques will follow.

2.1 Constant-wavelength (CW) neutron powder diffraction

To understand how reactor-based neutron diffraction (NPD) was born from a convergence of quantum mechanics and X-ray crystallography and became a practical tool during the Manhattan project, it is important to place the method development in the historical context. By the time neutrons were discovered by Chadwick in 1932, 12 both the De Broglie hypothesis of wave-particle duality had been confirmed and X-ray crystallography was well established already. In fact, the number of crystal structures reported by then based on X-ray diffraction (XRD) was in the hundreds. It was thus only a question of time before neutron diffraction by means of crystalline materials was proposed and demonstrated. Already in 1936, Elsasser considered neutron diffraction theoretically 13 and Preiswerk and von Halban, 14 and Mitchell and Powers 15 experimentally demonstrated that for polycrystalline iron and single-crystal MgO, respectively, neutron diffraction indeed occurred according to the Laue–Bragg conditions. However, the low intensities of the Ra–Be sources available at that time precluded any quantitative crystal structure analysis. The situation radically changed after nuclear reactors were built as a part of the Manhattan project. The first nuclear reactors CP-1 and X-10 went critical in 1942 and 1943, at what later became the Argonne and Oak Ridge National Laboratories, respectively. Both reactors produced high neutron fluxes such that neutron diffraction experiments began almost immediately, within months after the criticality was achieved. 16 , 17

Nuclear reactors generate a white beam with an approximately Maxwellian distribution of neutrons in energy or wavelength (Figure 2). To perform constant-wavelength (CW) neutron diffraction similarly to constant-wavelength X-ray diffraction, the beam should be coherent and collimated, and it must be monochromatized first. This can be achieved either with mechanical velocity selectors or with single crystal monochromators (Figure 5). In both cases, neutrons with a particular wavelength or energy are selected from the white beam as illustrated in Figure 2, by the shaded area. With velocity selectors the band position is controlled by the speed of rotation of neutron-absorbing blades. With monochromators, on the other side, it is controlled by the single-crystal angle in the incident beam via Bragg’s law (λ = 2d sin θ) for a fixed lattice plane with lattice spacing d hkl . As can be seen from Figure 2, only a small fraction of the neutrons produced by the reactor is utilized in CW neutron-diffraction experiments. While velocity selectors are still being used in the design of some instruments, e.g., for small-angle neutron scattering, single-crystal monochromators proved to be more suitable for neutron powder diffractometers, e.g., with respect to the desired instrumental resolution and the possibility to realize focusing optics gaining intensity. Conceptually, the two-axes design of CW neutron powder diffractometers has not changed since the instruments built at the first reactors. 18 The “two-axes” refer to one axis of rotation of single-crystal monochromator and another one as regards the detector around the sample, usually dubbed transmission Debye–Scherrer geometry. Nowadays position-sensitive detectors instead of point detectors are being used, making data acquisition far more efficient. A typical design using a “banana-like” detector is schematically illustrated in Figure 5.

Figure 5: 
Schematic sketch of a two-axis neutron diffractometer (left); α
1, α
2, and α
3 indicate beam divergence (typically defined by Soller collimators), β is the mosaic spread of the single-crystal monochromator. Scattering geometry (right) relating propagation vectors of the incoming and outgoing (diffracted) beams (k and k
0, respectively) and scattering angle of Bragg’s law (θ). See also Section 2.1.
Figure 5:

Schematic sketch of a two-axis neutron diffractometer (left); α 1, α 2, and α 3 indicate beam divergence (typically defined by Soller collimators), β is the mosaic spread of the single-crystal monochromator. Scattering geometry (right) relating propagation vectors of the incoming and outgoing (diffracted) beams (k and k 0, respectively) and scattering angle of Bragg’s law (θ). See also Section 2.1.

The monochromatic beam is sent onto the sample, and the distribution of the intensity of the beam scattered by the sample is measured as a function of the angle 2θ. Both positions and intensities of the characteristic diffraction peaks observed at particular 2θ (Bragg) angles contain information about the crystal structure and magnetic ordering of the sample material. The sharpness of the peaks in neutron powder diffraction (NPD) data is defined by the divergence of the beam, and the trade-off between resolution and intensity is controlled by the instrument optics. How the divergences α 1, α 2, α 3, the mosaic spread of the monochromator β, and the take-off angle 2θ M (Figure 5) affect the full widths at half maximum (FWHM) of the diffraction peaks and overall beam intensity (luminosity), was worked out by Caglioti and coworkers, 19 which thereby created a solid basis for the optimal design of CW neutron powder diffractometers. 20

Analysis of NPD data was originally performed by extracting structure factors from integrated intensities of the diffraction peaks, essentially the same procedure as for single crystals up to the present day. In fact, the very first crystal structures of NaH and NaD quantitatively studied by NPD were analysed that way. 21 Nonetheless, this very approach works only for simple data sets with well-resolved diffraction peaks. For complex structures and/or low-resolution diffraction data, peak overlap hinders extracting individual structure factors. To overcome this obstacle, the approach to powder diffraction data analysis was reversed and, instead, the diffraction profile is calculated based on the input structural model and then compared to the experimental data; for further improvement, the model gets iteratively adjusted to minimize the difference between model and experiment. This approach was developed in the late 1960s by Loopstra, van Laar, and Rietveld and is now commonly referred to as the Rietveld method. 22 , 23 , 24 , 25 Historically speaking, the conditions were favorable because of the nearly exactly Gaussian peak shapes of those diffractometers, e.g., in use at the neutron High Flux Reactor in Petten, The Netherlands, allowing for computing the so-called least-squares refinements by means of the upcoming computational power of that time. The method made tremendous impact and led to wide adoption of neutron powder diffraction, as well illustrated by the very rapid rise in the number of crystal structures reported based on neutron powder diffraction data from early 1970s (Figure 6). The Rietveld method has been implemented and further developed in dozens of software codes, and as of today the most widely used programs, e.g., FullProf, GSAS-II, TOPAS, Jana, Maud, can be used not only to refine crystal and magnetic structures but also extract a lot of other information as regards phase fraction, stress/strain, particle size, texture, extended defects such as stacking faults, and a lot more. 26

Figure 6: 
Number of the annual entries in the Inorganic Crystal Structure Database based on neutron powder diffraction data.
Figure 6:

Number of the annual entries in the Inorganic Crystal Structure Database based on neutron powder diffraction data.

Neutron powder diffraction combined with the Rietveld method proved to be so successful and in high demand that today most neutron scattering facilities have at least two neutron powder diffractometers: one optimized for high resolution aiming at clarifying complex crystal and magnetic structures and another one optimized for high intensity aiming at in situ or operando experiments on structurally known matter and small (down to a few mg) samples. Some specialized neutron diffractometers are also installed on “hot” and “cold” neutron sources with neutron beam spectra shifted to shorter and longer wavelengths (Figure 2), thereby focusing at short-range and long-period crystal and magnetic structures, respectively.

Theoretical work on optimal design, 20 decades of experience in experimentation, and in-depth knowledge exchange between neutron scattering facilities led to some common features in instruments. Most modern CW neutron diffractometers employ focusing Ge or graphite monochromators, providing wavelengths in the range of about 1.5–2.5 Å and 2.5–5 Å, respectively, 3He gas detectors, either multiple tubes or continuous multiwire tanks, and also interchangeable slits or Soller collimators 27 needed to adjust the resolution/intensity balance for different samples. To increase throughput, many diffractometers are equipped with automatic sample changers. As a result, data collection typically takes less than a few hours per data set and many CW NPD experiments are so streamlined that most neutron scattering facilities offer Mail-in (also known as Express or Rapid) access for which users post their samples to the facility and receive NPD data, collected at room and/or several low-temperature points, within a few days/weeks only. Also, high-temperature ovens and different kinds of high-pressure cells are usually routinely available. Given that there are about 40 constant-wavelength powder diffractometers around the world and data analysis with the Rietveld method is similar to the analysis of ubiquitous X-ray powder diffraction data, CW NPD certainly has the lowest entry barrier among all the neutron scattering techniques and should be routinely used by any research group working with crystalline materials, in particular when X-ray diffraction reaches its limits. Further information on the CW NPD can be found in ref . More particular resources for the instruments are found in Table 2.

Table 2:

Selected list of constant wavelength (CW) neutron powder diffraction (NPD) instruments

Facility CW NPD instrument
High resolution High intensity
Institut Laue-Langevin (France) D2B 33

www.ill.eu/users/instruments/instruments-list/d2b
D20 34

www.ill.eu/users/instruments/instruments-list/d20
Forschungsreaktor München II (Germany) SPODI 35

mlz-garching.de/spodi
ErwiN 36 (under construction)
Paul Scherrer Institute (Switzerland) HRPT 37

www.psi.ch/en/sinq/hrpt
DMC (cold neutrons)

www.psi.ch/en/sinq/dmc
Delft University of Technology (Netherlands) PEARL 38

www.tudelft.nl/en/faculty-of-applied-sciences/business/facilities/tu-delft-reactor-institute/research-instruments/pearl
High Flux Isotope Reactor (US) POWDER, HB-2A

neutrons.ornl.gov/powder
WAND2, HB-2C 39

neutrons.ornl.gov/wand
National Institute of Standards and Technology (US) BT-1 40

www.nist.gov/ncnr/high-resolution-powder-diffractometer-bt-1
Australian Nuclear Science and Technology Organisation (Australia) Echidna 41

www.ansto.gov.au/our-facilities/australian-centre-for-neutron-scattering/neutron-scattering-instruments/echidna-high
Wombat 42

www.ansto.gov.au/our-facilities/australian-centre-for-neutron-scattering/neutron-scattering-instruments/wombat-high
Japan Atomic Energy Agency JRR-3 reactor (Japan) HERMES 43

jrr3.jaea.go.jp/jrr3e/2/21.htm
China Advanced Research Reactor (China) HRPD 44 HIPD 44
China Mianyang Research Reactor (China) Xuanwu 45 Fenghuang 46
Korea Atomic Energy Research Institute (Korea) HRPD 47 HIPD 48
Bhabha Atomic

Research Centre (India)
PD-2, 49 PD-3 50 PD-1 51

2.2 Time-of-flight method and high-resolution neutron powder diffraction

To begin with, the striking advantage of the time-of-flight (TOF) method in neutron diffraction experiments is to utilize a wide range of the entire energy spectrum whereas the constant-wavelength angle-dispersive method uses only a part of the neutron spectrum by cutting out the rest, say, by using a monochromator (see Section 2.1). This was pointed out by Egelstaff (UK) in the 1950s already, and the intensity gain over the angle-dispersive method was discussed shortly after. 52 The TOF diffraction experiments were carried out for the first time using a so-called Fermi chopper at Swierk, Poland, 53 , 54 , 55 then using the pulsed reactor IBR-1 in collaboration with Dubna researchers. 56 Later on, the TOF method was continuously developed in electron accelerator facilities in which accelerated electron beams targeted to heavy metals generated neutrons through the (γ, n) reaction (see Section 1.1). Three facilities, Rensselaer linac (USA), Tohoku University linac (Japan), and Harwell linac (UK), independently carried out the TOF neutron diffraction and started fundamental research on the generation and scattering of pulsed neutrons. 57 , 58 , 59 , 60 In addition, a TOF Debye–Scherrer diffractometer with multiple detector banks was constructed by Kimura and coworkers 58 , 59 to conduct crystal and magnetic diffraction experiments on metals and oxides.

While many pulsed-neutron source projects were proposed but not funded, spallation neutron projects using proton accelerators opened the next era. Four spallation facilities first opened user programs since the 1980s: IPNS (1981–2008, Argonne National Laboratory, USA), KENS (1980–2006, High Energy Accelerator Research Organization, Japan), LANSCE (from 1983 until now, Los Alamos National Laboratory, USA), and ISIS (from 1984 until now, Rutherford Appleton Laboratory, UK).

The recent important advance in TOF diffraction comes from the worldwide construction of high-intensity proton accelerators: ISIS (UK), SNS (from 2007 until now, Oak Ridge National Laboratory, USA), J-PARC MLF (from 2008 until now, Japan Atomic Energy Agency and High Energy Accelerator Research Organization, Japan), China Spallation Neutron Source (from 2018 until now, Institute of High Energy Physics, China) and the coming European Spallation Source (ESS, Sweden). The many scientific publications of work based on measurements at these large-scale research facilities provide a fine testimony to the success of the TOF neutron-diffraction technique.

In spallation, pulsed protons colliding with a heavy metal target effectively produce high-energy pulsed neutrons (see Section 1.1), which are then moderated to energies in the range of 0.1 meV–100 eV using a moderator. This energy range corresponds to the wavelength range of 0.03–30 Å by the de Broglie relationship, λ = h m v = h 2 m E , as previously defined in Section 1.1. Although neutrons with various energies (wavelengths, velocities) are generated at the same time, the time when the neutrons are scattered by the sample and finally reach the detectors is different, thereby reflecting the difference in neutron velocity.

The TOF method measures the elapsed time for a neutron from a defined state with time t 0 = 0, e.g., after the spallation impact and slowing down at the moderator or by passing a chopper window creating a pulsed beam, until the time where the neutron gets scattered by the sample, and finally reaches the detector. We may therefore say that each neutron event (release, propagation, scattering, and detection) can be distinguished and recorded in a neutron scattering experiment. Typically, 106–108 neutron events are accumulated for a crystal structure analysis and a time channel on the order of 104 μs. The neutron-event data are usually converted to histograms with linearly or logarithmically binned time channels of a few μs intervals for further data analysis such as Rietveld refinement.

Since the neutron velocity, v, does not change for the elastic Bragg scattering, v can be calculated by v = L t in TOF diffraction (L: total flight path, t: elapsed time). Combining this equation with the de Broglie relationship, we get λ = h t m L . By using the Bragg condition, λ = 2dsinθ (d: lattice spacing, 2θ: scattering angle), we obtain

(8) t = 2 m L sin θ h d ,

which shows that the time of arrival of the neutron at the detector is proportional to the d spacing (see Figure 7), an exceptionally simple and beautiful relationship.

As given by Figure 7, it appears that slower and longer-wavelength neutrons belonging to one neutron pulse arrive at the detectors after the newly released neutrons (from a later neutron pulse but with faster and shorter wavelengths) have also arrived. This unwanted phenomenon is called frame overlap, so the measurable wavelength range is limited by the frame overlap. There are several methods to extend the limit of the measurable d range due to frame overlap. As can be seen from equation (8), a Bragg reflection with a larger d can be observed using lower-angle detector banks. The measurable d range is often extended by lowering frequencies using mechanical choppers; multiple disk choppers in lower frequencies remove faster neutrons and a t 0 chopper removes burst neutrons with very high energies which otherwise contribute to the background.

Figure 7: 
A conceptual diagram in TOF diffractometry with a pulsed source and 25 Hz frequency. The Bragg reflection positions, t
1, t
2, …, are a function of d, flight path, L1 + L2, and a scattering angle, 2θ.
Figure 7:

A conceptual diagram in TOF diffractometry with a pulsed source and 25 Hz frequency. The Bragg reflection positions, t 1, t 2, …, are a function of d, flight path, L1 + L2, and a scattering angle, 2θ.

As many neutron-diffraction practitioners do, it is convenient to remember the following formula:

(9) d [ Å ] = t [ μ s ] 505.555 [ μ s Å m ] L [ m ] sin θ

to achieve the best diffraction strategy.

Although the advent of the Rietveld method allows extraction of structural information from overlapping Bragg reflections, a too severe overlap makes the analysis results unreliable, especially for very small lattice spacings. In contrast, the use of high-resolution diffractometers with a narrow Bragg peak full width at half maximum (FWHM), ∆d, may lead to new discoveries because the diffraction patterns contain a plethora of information about the structure.

In the early days of TOF diffraction, the TOF technique was not thought to be suited to high-resolution powder diffraction, 20 interestingly enough. In fact, the 150 m flight-path high-resolution diffractometer with the resolution of 2 × 10−3 needed a large amount of sample and measurement time. 61 However, due to technological developments in neutron transportation, detection, diffractometer design, sources, etc., the TOF method is very well suited to high-resolution powder diffraction, even though the shape of the Bragg peaks appears as being a little more complex than in the constant-wavelength scenario.

It is reasonable to define the resolution, d d , of the TOF powder diffractometer as composed of three variances: t t , L L and cotθΔθ. To achieve high resolution, these three variances should be small, that is, a large L, which additionally translates into a large t, and a large 2θ.

High-resolution TOF powder diffractometers with the best resolution of ca. d d 5 × 10−4 such as HRPD at ISIS, SuperHRPD at J-PARC, and HRPD at CSNS have long flight paths, about L = 80–100 m (both L L and t t are small) and backward detector banks (cotθΔθ is close to zero).

It should be emphasized that the latest high-resolution hydrogen moderator cooled to about 20 K are designed to have sharp neutron pulses with small ∆t over a wide energy range from 1 meV to 300 meV (0.5 Å to 10 Å).

Finally, Table 3 gives a list of general-purpose high-intensity and high-resolution TOF neutron diffractometers as well as instruments for special science cases at a world-wide selection of neutron sources. Future TOF instruments are currently built at the ESS in Sweden (e.g., DREAM, HEIMDAL) and at the FRM-II/MLZ in Germany (POWTEX).

Table 3:

Selected list of time-of-flight (TOF) neutron powder diffraction (NPD) instruments.

Facility TOF neutron powder diffractometers
High-resolution diffractometers High-intensity diffractometers Special diffractometers
ISIS (UK) HRPD

https://www.isis.stfc.ac.uk/Pages/

hrpd.aspx

WISH

https://www.isis.stfc.ac.uk/Pages/

wish.aspx
GEM

https://www.isis.stfc.ac.uk/Pages/

gem.aspx

Polaris

https://www.isis.stfc.ac.uk/Pages/

polaris.aspx
PEARL: High Pressure Diffraction

https://www.isis.stfc.ac.uk/Pages/

pearl.aspx

ENGIN-X: Engineering Diffraction

https://www.isis.stfc.ac.uk/Pages/

engin-x.aspx
SNS (US) POWGEN

https://neutrons.ornl.gov/powgen
NOMAD

https://neutrons.ornl.gov/nomad
SNAP: High Pressure Diffraction

https://neutrons.ornl.gov/snap

VULCAN: Engineering Diffraction

https://neutrons.ornl.gov/vulcan
LANSCE (US) SMARTS:

High Pressure & Preferred Orientation

https://lansce.lanl.gov/facilities/lujan/instruments/smarts/index.php

HIPPO:

Engineering Diffraction

https://lansce.lanl.gov/facilities/lujan/instruments/hippo/index.php
J-PARC (Japan) SuperHRPD

https://mlfinfo.jp/en/bl08/

iMATERIA

https://mlfinfo.jp/en/bl20/

SPICA

https://mlfinfo.jp/en/bl09/
NOVA

https://mlfinfo.jp/en/bl21/
PLANET: High Pressure Diffraction

https://mlfinfo.jp/en/bl11/

TAKUMI: Engineering Diffraction

https://mlfinfo.jp/en/bl19/
CSNS (China) TREND

(under construction)

http://english.ihep.cas.cn/csns/fa/in/202109/t20210915_283265.html

GPPD

http://english.ihep.cas.cn/csns/fa/in/202109/t20210910_283134.html
MPI

http://english.ihep.cas.cn/csns/fa/in/202109/t20210915_283259.html
High Pressure Diffractometer

(under construction)

http://english.ihep.cas.cn/csns/fa/in/202109/t20210915_283264.html

EMD: Engineering Diffraction

http://english.ihep.cas.cn/csns/fa/in/202109/t20210915_283266.html

2.3 Multidimensional Rietveld

We recall that constant-wavelength (CW) and time-of-flight (TOF) diffraction both probe the lattice spacing d according to Bragg’s law. While monochromatic CW diffractometers screen the scattering angle 2θ with their detector, a very simple TOF diffractometer concept would place one detector at a fixed scattering angle while the wavelength λ varies within a neutron pulse. In modern neutron TOF diffractometers such as POWGEN, SNS (USA), the detector system covers almost the entire range of scattering angles, i.e., both 2θ and λ vary. Every detector pixel at fixed 2θ measures an individual TOF pattern, and each d reflection is “moving” over the detectors with λ and within the selected wavelength band. For sticking to a one-dimensional, conventional data-treatment, the detectors are grouped into detector banks with narrower 2θ ranges in which the data are integrated along d to yield (several) TOF patterns with similar resolution Δd/d.

The sophisticated detector arrangement of POWGEN 62 allows to integrate the whole detector (one bank) into one d-dependent pattern which is then converted back to a TOF pattern using Bragg’s law for 2θ = 90°. This procedure is called diffraction focusing (see Figure 8, left). As a clear advantage, these reduced pattern(s) – one per wavelength-band – can be processed with all common Rietveld packages such as FullProf, 65 GSAS-II, 66 etc., either as single patterns each or simultaneously in a multi-pattern refinement of several wavelength-bands. However, for every instrumental setting (combinations of detector bank, wavelength range, chopper settings, etc.) an individual, averaged instrument parametrization is needed. This is because the resolution, i.e., the reflection-width Δd, is averaged during diffraction focusing. Similarly, the reflection shape, being convolutions of Gaussian, Lorentzian and other functions with higher complexity (see Section 2), and its parametrization is changing with (and within) each setting. Both effects may compromise the maximum instrument performance, so a loss of information is likely.

Figure 8: 
Left: Diffraction focusing from an angular- and wavelength dispersive pattern to a conventional TOF pattern based on a CuNCN sample
63

,

64
 measured at POWGEN. Right: Schematic drawing showing one arbitrarily chosen d
⊥ (red curve) in (2θ, λ) space being always normal to all d-reflections and the comparison to CW (blue, horizontal cut) and TOF (green, vertical cut). Adapted from ref 10 by CC BY.
Figure 8:

Left: Diffraction focusing from an angular- and wavelength dispersive pattern to a conventional TOF pattern based on a CuNCN sample 63 , 64 measured at POWGEN. Right: Schematic drawing showing one arbitrarily chosen d (red curve) in (2θ, λ) space being always normal to all d-reflections and the comparison to CW (blue, horizontal cut) and TOF (green, vertical cut). Adapted from ref 10 by CC BY.

For the multi-dimensional Rietveld method, the angular- and wavelength-dispersive neutron events are binned without diffraction focusing either in a (2θ, λ) coordinate system (see Figure 8, right) or in the recently introduced (d, d ) coordinate system with d as d-spacing from (2θ, λ) and

(10) d = ( λ 2 2 λ K 2 ln ( cos θ ) ) 1 / 2  with  λ K 2 = 1 Å 2

being the solution to the differential equation dλ/dθ = –(2d cos θ)−1. The differential equation requests d to be perpendicular to all d-reflections. 10 Accordingly, the sinusoidal d-reflection in (2θ, λ) becomes a perfectly vertical line. By definition, the d axis will always be perpendicular to the d-reflection in both, (2θ, λ) and (d, d ), coordinate systems (see Figure 8, right). Furthermore, it will intersect the d-reflection with the smallest possible reflection-width Δd, being the instrumental resolution for an ideal sample. Consequently, in the multi-dimensional description, the instrumental resolution function depends on 2θ and λ. Such a function, based on the fundamental instrument characteristics (divergence of neutron beam, sample size, pixel size, pulse-width, etc.) of POWGEN, has been fully described. 67

Instruments such as GEM, POLARIS, at ISIS and SuperHRPD at JPARC, as well as the future instruments POWTEX, FRM II (Germany) and DREAM, ESS (Sweden) have a very large solid-angle detector coverage. The detector geometry is very simple as the detector-surface is cylinder-like sharing its axis with the neutron beam axis, 68 different from the heart-shaped layout at POWGEN. Hereby, the instrumental resolution function is changing even more drastically and makes the multi-dimensional description more favorable. To test the tailor-made POWTEX detectors and develop the multi-dimensional Rietveld method, one mounting-unit (Δ2θ = 90°; Δφ ≲ 9°) plus electronics was operated at POWGEN with diffraction conditions very similar to POWTEX’s secondary instrument. 69 The detector was calibrated and the measured event data were treated using the PowderReduceP2D 70 algorithm as implemented in Mantid 71 to yield the binned diffraction patterns in (2θ, λ) and (d, d ) coordinates, given the fundamental instrument resolution function as defined similar to ref 72, and all intended to test the multi-dimensional Rietveld method implemented in a modified, yet unpublished GSAS-II version.

Powdered diamond with very sharp reflections serves as an excellent example for illustration. By refining the parameters of the structural model (unit cell, isotropic thermal displacement, background, instrumental parameters) to make the calculated pattern eventually fit the experimental neutron diffraction pattern, the multidimensional fit (see Figure 9) is already convincing by visual inspection. A real-world sample of BaZn(NCN)2, however, not only evidences similar standard deviations compared to the conventional approach but arrives at internal structural parameters which are slightly different and make more chemical sense. For example, two formerly very similar lattice parameters (a and b) of the orthorhombic structure are better resolved (and differ more), and this also relates to even finer structural details as regards C=N double-bond lengths and N–C–N bond angles. 69 So, the quality of structural information, i.e., the related scientific question, clearly profits from the multidimensional approach.

Figure 9: 
Multi-dimensional Rietveld refinement of a powdered diamond sample. Top, from left to right: observed, calculated and difference pattern. Bottom: 1D plot of multi-dimensional refinement (left) and conventional, one-dimensional refinement using an unmodified GSAS-II (right). Blue lines indicate peak positions. Adapted from ref 69 by CC BY.
Figure 9:

Multi-dimensional Rietveld refinement of a powdered diamond sample. Top, from left to right: observed, calculated and difference pattern. Bottom: 1D plot of multi-dimensional refinement (left) and conventional, one-dimensional refinement using an unmodified GSAS-II (right). Blue lines indicate peak positions. Adapted from ref 69 by CC BY.

2.4 Texture

An inherent feature of most polycrystalline materials and aggregates, be they natural (like rocks) or synthetically produced (like metals and ceramics), is the preferred orientation of the constituting crystallites. This preferred orientation distribution of the crystallites is called the crystallographic texture. 72 Because of the anisotropic properties of most single crystals, texture influences mechanical and physical properties such as strength, conductivity, magnetic susceptibility, light refraction, or wave propagation in materials. Deformation, casting, welding, and heat input may change the texture and, thus, the knowledge of the texture and its evolution is a key issue in property optimization in materials science. 73 In addition, texture provides a wealth of information on geological processes and the manufacturing routes of archaeological artefacts. 74

The ultimate aim of texture analysis is to quantitatively obtain the three-dimensional (3D) orientation distribution of the crystallite grains in a polycrystalline material with respect to a sample coordinate system. Besides the shape this is often defined in terms of directions of the forming work acting on the sample, like the rolling direction in metals after cold processing or foliation in rocks. In general, the resulting orientation distribution function (ODF) can be given as the volume distribution of grains dV in an orientation increment dg defined by the Euler angles.

In diffraction the integral intensity of the scattered beam is proportional to the volume fraction of crystallites within the sample volume fulfilling the Bragg condition. By rotating the sample through all possible orientations, the intensity distribution across a polar sphere can be measured. In a diffractometric texture measurement this is usually mapped to a stereographic projection and yields the so-called pole figure. Manipulation of the sample orientation is carried out using Eulerian cradles or, more recently, robots. 75 , 76 As the information on the rotation of crystallites around the scattering vector is lost in such a measurement, a pole figure only represents a 2D projection of the true ODF. Therefore, several pole figures are needed to calculate the ODF using pole figure inversion. 72 , 77

Since texture analysis using neutron diffraction was first used by Brockhouse in 1953 in the attempt to determine the magnetic structure of nickel, 77 it has seen a wide range of applications ranging from the earth sciences 78 to applied industrial processing technologies. 73 Texture analysis can be carried out on neutron powder and single-crystal diffraction data; hence it is applicable at all neutron sources, that is, reactor, spallation and accelerator-based. In contrast to other methods, neutron diffraction usually measures complete pole figures. 79 Due to the weak interaction of neutrons with matter, effects of absorption are almost negligible and large sample volumes in the order of several cm3 can be probed (see Section 1.2). This ensures an accuracy better by almost one order of magnitude than by using X-rays, simply due to the large number of sampled grains, but it also allows for coarse grained materials to be investigated. Non-destructive bulk texture analysis needs no special sample preparation such that neutron diffraction has shown its enormous potential as an efficient tool for measuring multi-phased materials. Using TOF diffractometers with multi-detector systems adds the benefit that many Bragg reflections can be recorded simultaneously, and fewer sample rotations are necessary for fast texture measurements. 80 Especially for crystallites of low symmetry with many closely spaced or even overlapping reflections (a very typical scenario found for many rocks or similar geological samples), TOF neutron diffraction in combination with advanced data analysis methods like Rietveld texture analysis is essential. 81 Finally, neutron diffraction uniquely offers the possibility to study the magnetic texture of polycrystalline materials.

2.5 Single-crystal neutron diffraction

As mentioned before, powder diffraction is one of the most successful and fastest methods to obtain structural information on the atomic scale. The applications are numerous and cover all the important fields of research from solid-state physics, chemistry, materials science and mineralogy, as well as biology and macromolecular sciences. Nevertheless, diffraction on single crystals has the ability to contribute additional details about the spatial distribution of the various intensities from differently orientated lattice planes. A highly schematic picture of both powder and single-crystal diffraction in presented in Figure 10. The latter method allows to record the specific intensities of different lattice planes separately, even if their d values and respective |Q| values are the same or so similar that they overlap in powder diffraction experiments. Examples where the spatially resolved assignment of intensities to individual lattice planes becomes particularly advantageous are given below.

Figure 10: 
Constant wavelength diffraction patterns for powder (left) showing Debye–Scherrer rings per d lattice-spacings and single crystal (right) showing distinct spots; in the single crystal case the sample needs to be rotated in order to fulfill Bragg’s law for one reflection after each other; in the ideal powder sample a random distribution of lattice-planes is naturally present.
Figure 10:

Constant wavelength diffraction patterns for powder (left) showing Debye–Scherrer rings per d lattice-spacings and single crystal (right) showing distinct spots; in the single crystal case the sample needs to be rotated in order to fulfill Bragg’s law for one reflection after each other; in the ideal powder sample a random distribution of lattice-planes is naturally present.

Many applications benefit from combining the advantage of spatial information preservation through single crystal diffraction with the excellent discrimination between Q-dependent and Q-independent neutron interactions. One example are the mean-square displacements (MSD) – also going under the name atomic displacement parameters (ADP) in the Debye–Waller part of the structure factors – whose temperature dependence can reveal information about thermal oscillations and static positional disorder. Access to large Q with neutrons allows for very accurate studies that might help to carefully validate density-functional based phonon simulation studies. 82

Another example is the Q-dependent magnetic scattering length respective form factor (see Section 1.2). The direction-dependent observation of the magnetic and non-magnetic intensity components of individual lattice planes allows for very precise studies on the structural character of macroscopic order phenomena (ferro-, ferri- or antiferromagnetism) and the orientation of the magnetic moments, in particular as regards potentially important materials for information technology. 83 By spin-polarizing the magnetic moment of neutrons in the beam, this allows for even more detailed recording of the vectoral components of magnetic elements (e.g., 3D polarimetry, spin-flip-ratio measurements), which cannot be discussed in greater detail within the scope of this primer, 84 , 85 but please see Section 3.4.

Resolving twinned structures is another domain of single-crystal diffraction needed to clarify effects of crystallographic symmetry lowering. The latter is a consequence of crystallographic transformation by group-subgroups relationships, 86 , 87 accompanied by domain formations often caused by small displacements of the light elements; on the atomic scale, the higher symmetry has gone but macroscopically it is (almost) retained. Typically, a reduction in symmetry is accompanied by additional reflections (so-called superstructure reflections). If they are caused by small displacements of light elements they can easily be missed upon using X-ray diffraction. Furthermore, twinning causes the creation of domains in a single crystal whose spatial orientations to each other follow specific so-called twin laws. 88 The overall intensity distribution of a twinned single crystal is thus the result of a complicated – often only partial – overlap of the intensity patterns of the different domains as well as their volumetric fractions and may also include reflection splitting (for pseudo-merohedral twins). In powder diffraction, intensities are plotted against d-spacing and |Q| respectively. Additional spatial information like the angular relationships between equivalent lattice planes of different domains can be accurately recorded with single crystal diffraction. This makes it much easier to correctly assign the measured intensities (including magnetic ones) to the intensity patterns of the different domains and thereby to identify their true space group. 83

The reader may recall the main advantages of diffraction with neutrons covered in Section 1.2, that is, the comparatively strong scattering lengths of many light elements (H, Li, O, N, etc.) and the fact that they are independent from the scattering vector Q = kk 0, the difference between outgoing and incoming wave vectors k and k 0, respectively, also dubbed k f and k i in the literature (see Figure 5, right). The length of the vector Q relates to the lattice-spacing by d = 2 π | Q | . Hence, single-crystal neutron diffraction, in particular, may localize those atoms precisely, for instance in molecules or in ion conductors for energy applications 89 or minerals for environmentally friendly applications. 90 , 91 As also covered before, a keener look reveals that NPD suffers from strong incoherent scattering in the case of hydrogen which leads to an enhanced background and to a massive deterioration in the signal-to-noise ratio. While this problem may be solved by exchanging hydrogen by deuterium since deuterium often behaves chemically very similarly to protium (see Section 1.2), such approach is not always applicable, for instance, when studying natural minerals. Furthermore, exchanging isotopes may significantly impact the chemical behavior, e.g., for hydrogen-bridging bonds, or affect the crystal structure dynamics due to the higher mass of D compared to H such that protium-deuterium replacement may shift the temperatures of phase transitions or even generate new phase transitions. 92 Single-crystal diffraction, however, suffers much less from this issue of incoherent hydrogen scattering, as the diffracted Bragg intensities concentrate on well located small areas in reciprocal space, so they are not distributed over the full Debye–Scherrer rings. In many cases, single-crystal neutron diffraction preserves a good signal-to-noise ratio even for non-deuterated samples and allows to solve the structures of small non-deuterated molecules. 82 , 93

Furthermore, the nuclear scattering lengths of neighboring atoms often differ considerably (e.g., Mn vs. Fe, see Section 1.2 and Figure 3). This is an interesting feature for studies on mixed crystals, where X-ray diffraction often fails to discriminate the site distribution (and occupancies) of elements with similar electron configurations. 94 , 95 Instead, using the different scattering lengths of various isotopes of the same element (like in the H/D exchange) can improve contrast variation. At the same time, the absorption cross sections of most elements in the periodic table are very small and allow neglecting absorption correction even in large single crystals of few millimeter size, very much different from single-crystal X-ray diffraction. There is another effect dubbed extinction affecting intensities and going back to microstructural features which also requires different approaches for correction in X-ray and neutron scattering.

Of course, neutron single-crystal diffraction does not only offer advantages; there is also a price to be paid. To begin with, neutron sources deliver significantly lower neutron fluxes compared to the high photon fluxes of synchrotron radiation sources. At the same time, the interaction probability of neutrons is much lower (nuclear instead of electron scattering, see Section 1.2). In order to preserve reasonable counting statistics for neutron experiments, they are performed with a worse resolution (Δd/d ratio) from the source/monochromator and much larger sample volumes (millimeter sized instead of micrometer sized) compared to the X-ray counterparts. Thus, growing suitably sized single crystals for neutron single-crystal diffraction can be very challenging. Even then, measurements at neutron facilities require significantly more time than at synchrotron facilities.

The larger sample volumes also affect single-crystal-typical effects such as multiple scattering at the same (dubbed extinction effect, see above) or different (dubbed Renninger effect) sets of lattice planes. Both effects do also exist for X-ray diffraction, but special attention must be paid to extinction in order to avoid misinterpretations, especially for very good and large crystals, while the Renninger effect happens more often due to the poorer energy resolution.

We close this section with a few remarks concerning instrumentation: most neutron single-crystal diffraction experiments at reactor (= nuclear fission) sources are carried out in angular dispersive mode which requires a monochromatic beam. By rotating the sample, the intensities of one set of lattice planes are collected one after each other, putting the detector on the corresponding Bragg angle (or using an area detector). Nonetheless, the energy-dependent velocities of neutrons also allow for the so-called time-of-flight method (see Section 2.2), an energy dispersive technique that prospers from the evolution of spallation sources which generate high peak intensities from a pulsed neutron flux. The fixed single crystal sample is then surrounded by a large area detector. Combining the positions and times of arrival of the diffracted neutrons on the detector reveal the corresponding lattice planes in a comparatively short period of time. This makes the technique most suitable for experiments with limited movability of the sample, for instance experiments at temperatures in the milli-Kelvin range and within strong magnets.

Recently, considerable effort has been made to investigate diffuse scattering with neutrons. Further information can be found in Section 3.1 or, e.g., in ref 96.

3 Specific applications

3.1 Diffuse scattering

Generally speaking, all diffraction data include both the information from the average structure and from any deviations from the average lattice or spin structure. The latter part is called diffuse scattering – see ref 97 for a readable historical overview – and is due to spatial fluctuations in occupation and atomic positions and the respective magnetic moments. The isotropic background of Laue scattering 98 from a binary solid solution provides the simplest example. The truly ideal case is found in the incoherent scattering of isotopic mixtures or randomness of nuclei with spin. For real solid solutions, like ZrTi, a modulation of the Laue intensity is observed indicating a deviation from simple randomness, such as preferential near-neighbor pairing. Diffuse scattering often shows very characteristic features, serving as fingerprints of the underlying defect structure and short-range correlations that impact material properties, e.g., the short-range order in alloys, which appears as precursor states near ordering phase transitions. The diffuse scattering from single crystals between the Bragg peaks is typically very weak in intensity, when compared to Bragg scattering, unless integrated over the full solid angle. Accordingly, modern neutron instrumentation tends to use large area detectors, which is obviously most efficient for measuring diffuse scattering.

There are several approaches to reveal the microscopic origin of diffuse scattering, either by Fourier analysis (of short-range order) in terms of pair correlation functions, 99 or by modeling with a certain assumed motif, 100 or by trying to establish a real-space model by the reverse Monte Carlo method 101 , 102 and, similarly, in the case of spin structures by the Spinvert software. 103 This approach is limited to simple cases, however, because the intensity is a sum of possibly indistinguishable terms from many pair correlations in multi-component systems. An alternate approach, PDF analysis (see Section 3.2), provides model-free insights into real space as given by the superposition of the pair correlation function.

Some neutron instruments are dedicated to diffuse scattering studies, e.g., the SXD – the single crystal diffractometer at the ISIS spallation source. 104 A large volume in reciprocal space is explored by neutron time-of-flight Laue diffraction in a single setting with an area detector using a pulsed, polychromatic beam. The instrument CORELLI at the spallation neutron source SNS in Oak Ridge uses a pseudo-statistical chopper for an efficient separation of elastic diffuse scattering from structural disorder, while discriminating thermal inelastic background. 105

3.2 Pair distribution function analysis

Whenever structures deviate from perfect crystallinity such as in nanostructured materials, disorder evolves and gives rise to diffuse scattering (see section before). Common powder diffraction experiments relying on the interpretation of only the Bragg peaks therefore face limitations in the structural characterization of such disordered materials, liquids, and glasses. To then achieve more insight, one needs to introduce the pair distribution function (PDF) which is the probability of finding two atoms in the sample separated by distance r, weighted by their scattering lengths b i . The PDF is accessed by Fourier transformation of diffraction data collected to high momentum transfer Q. 106 Hereby, we use not only scattering from Bragg peaks, but additionally any diffuse scattering in-between and underneath the Bragg peaks, resulting in the terminology total scattering. It is important to note that the mathematical definitions of radial and pair distribution functions vary slightly between different publications, very unfortunately, and that G(r) is commonly used for the X-ray PDF while for neutron PDF (nPDF) further definitions co-exist employing D(r) or T(r). Please see ref 102 for a comparison of nomenclatures.

Originally, the short-range order and hydrogen bonding networks in bulk liquids 107 and around ions in solution 108 have been commonly investigated, as well as the structure of glasses. 109 The general popularity of PDF analysis exploded with the need for structural characterization of nanomaterials 110 often lacking sufficient crystallinity. For clusters of a few atoms, quantum dots, nanocrystals, but also bulk materials with nanoscale domains, the PDF approach will excel in extracting the interatomic distances, coordination numbers, local vibrational motion, local structural motifs and chemical short-range order in the sample, as well as nanoparticle size. Because of the constant, Q-independent neutron scattering lengths, in contrast to the decaying X-ray atomic form factor, nPDF studies can measure to high momentum transfer Q max, readily up to 50 Å−1. Hence, smaller changes in interatomic distances are accessible simply because of better resolution in the PDF, as defined by Δr = 2π/Q max.

The collection of PDF data is based on a powder diffraction experiment up to high Q of at least ≈ 20 Å−1. To achieve the latter, short neutron wavelengths are employed, often combined with extensive detector coverage to collect the entire angular range simultaneously. State-of-the art nPDF instruments are found at spallation sources enabling high flux at high neutron energies, i.e., at CSNS, ISIS, SNS, and JSNS. At present, only the hot source of the ILL features a reactor-based PDF instrument. A very recent review on nPDF recapitulates facilities, data treatment, and analysis with examples. 111

One recent nanomaterials example showcasing the power of nPDF analysis was the revelation of a much higher concentration of surface oxygen defects in CeO2 rod-shaped nanocrystals compared to ceria nanocubes. The refinement of the nPDF data distinguished surface from bulk oxygen defects and required two different phases, one of bulk CeO2 with Frenkel defects and one of surface Ce3O5+x species, see Figure 11. Further, a smaller number of surface defects has been observed upon exposure of the nanorods to SO2 being possibly the reason for sulfur poisoning of these catalytic materials. 112 The high sensitivity of neutrons towards light elements such as H, C, O and Li is exploited in nPDF studies of the local structure of fuel cells, heterogeneous catalysts, ferroelectrics, and battery materials. For instance, nPDF helped to clarify that an increase in Co concentration in proton-conducting oxides of the type La0.9Sr0.1Sc1−x Co x O3−δ impacts the local structural disorder and proton location. 113 This field of energy materials comprises the need for in situ and operando measurements (see Section 3.5).

Figure 11: 
Left: Neutron PDF refinement of ceria nanorods with two phases: ceria with Frenkel-type defects by applying a numerical nanorod shape correction γ
0(r) and surface defects modelled as Ce3O5+x. The overall fit (red) to the experimental data (black) results in the difference curve (blue, in offset), showing the shape envelope function γ
0(r) (purple), as well as the PDF of the contribution of surface defects (green, in offset). Right: Scheme of the surface species Ce3O5+x as spherical nanoparticles on the surface of the ceria nanorods together with their crystal structures. Adapted from ref 112 by kind permission of the American Chemical Society.
Figure 11:

Left: Neutron PDF refinement of ceria nanorods with two phases: ceria with Frenkel-type defects by applying a numerical nanorod shape correction γ 0(r) and surface defects modelled as Ce3O5+x. The overall fit (red) to the experimental data (black) results in the difference curve (blue, in offset), showing the shape envelope function γ 0(r) (purple), as well as the PDF of the contribution of surface defects (green, in offset). Right: Scheme of the surface species Ce3O5+x as spherical nanoparticles on the surface of the ceria nanorods together with their crystal structures. Adapted from ref 112 by kind permission of the American Chemical Society.

The pair-correlation function, i.e., the Fourier transform of the scattering data from reciprocal to real space also provides a valuable approach to the analysis of magnetic diffuse scattering. Magnetic short-range order and correlations can efficiently be determined by acquiring diffraction data and utilizing the PDF formalism. 106 For powder diffraction, this magnetic PDF (mPDF) analysis 114 proceeds by modelling the structural pair correlation function which is then subtracted from the experimental pair correlation function. The difference then serves as a first model for the magnetic pair correlation such that subsequent refinement techniques can be used to iteratively model both the structural and magnetic contributions. The mPDF approach was utilized recently to model the short-range order in MnO 115 and frustrated magnets. 116 Likewise, for single crystals, the 3D-mPDF method can be utilized to determine the three-dimensional magnetic correlations. The analysis proceeds by removing both the structural and magnetic Bragg peaks, for example through the punch-and-fill or KAREN method, and then employs a 3D-Fourier transform of the remaining diffuse magnetic scattering to obtain the magnetic correlations in 3D. This method was recently showcased on MnTe single-crystal data to determine the real-space correlation in the short-range ordered state of this antiferromagnet. 117 We expect that PDF and mPDF analysis may see further advances and progress when combined with use of polarized neutrons, and when appropriate, dedicated instrumentation will become available for users.

3.3 Polarized neutrons

For revealing magnetic diffuse scattering, it is often helpful to use polarized neutrons as probes and possibly polarization analysis on the scattered neutrons. Polarized neutron diffraction has been established in the 1960s already. 118 , 119 Clearly, knowing about the direction and change of the neutron’s magnetic moment upon scattering will provide more detailed structural information. Furthermore, the higher quality data may justify using the lower flux of polarized experiments. This enables clean separation of magnetic diffuse scattering and elimination of any background, e.g., through the so-called XYZ analysis, applicable within the orientational average for powder diffraction using multi-detectors. 120 With polarized neutrons and polarization analysis it is possible to distinguish nuclear scattering from magnetic scattering, to probe sensitively the orientation of magnetic moments, and further to reveal what remains hidden to unpolarized studies, that is, the interference of nuclear and magnetic amplitudes, and the handedness (also dubbed chirality) in magnetic configurations. Single crystals and polarized neutrons are necessary requisites to directly reveal chiral properties by magnetic diffuse scattering from disordered spin structures. 84 , 121

Polarized neutrons are valuable to enhance the intensity signal |N ± PM | 2 for very small moments by interference of nuclear and magnetic amplitudes, N and M , as obtained from the intensity difference under reversal of the neutron beam polarization P. Note that the ⊥ index in M refers to only the components of M perpendicular to Q, which are visible to neutrons. 122 The magnetic structure can be established in a two-step process by first determining the nuclear, non-magnetic crystal structure. Moreover, the linear response of M to an applied magnetic field H is proportional to the susceptibility and can be obtained at the atomic scale by crystallographic analysis of the Bragg intensities. This analysis includes preferred orientations, anisotropies, and spin- and magnetization densities, which can be extracted efficiently even from powder data using 2D detectors 123 , 124 and from magnetic nanoparticles. 125 One can readily extend this approach to ferrimagnetic samples and determine weak magnetic amplitudes from small magnetic moments, even with high precision. Similar to the electron density distribution obtained from X-ray Bragg intensities, magnetic form factors and spin densities can be determined from magnetic Bragg intensities in neutron diffraction.

It is noteworthy that using polarized neutrons and polarization analysis, one can also distinguish spin-incoherent from coherent scattering, a possibility that may be of interest and relevant particularly for hydrogenous materials. However, the need for more specialized instrumentation and lower flux still limits applications.

Combining polarization analysis with Bragg and diffuse scattering will be possible at the new ESS diffractometers, 126 the polarized instrument MAGIC and in further future the instrument DREAM. 127 These will enable new types of experiments combining atomic and magnetic PDFs, for which just recently proof-of-principle experiments were carried out on MnTe at HYSPEC, SNS. 128

3.4 Magnetism

The first pioneering studies of determining magnetic structures 129 were carried out by analyzing powder-diffraction data using unpolarized neutrons, similar in style to regular structural refinements covered in Section 2.1, and this most frequent approach persists until today. 22 , 65 The magnetic structure information is found by the magnetic cross-section which imposes that only magnetization perpendicular to the scattering vector Q is observed. Moments parallel to the scattering vector are not observed, in principle. For layered materials in particular, magnetic intensity for Bragg peaks perpendicular to the layers vanishes for purely axial moments, see Figure 12, left. For antiferromagnets, in many cases the magnetic Bragg peaks do not overlap with structural peaks (see below, however), witnessing a new periodicity popping up from the diffractogram. For ferromagnets, however, magnetic Bragg peaks are superimposed on top of some nuclear peaks. The simplest approach to identify the magnetic peaks is thus to measure the diffraction pattern above and below the ordering temperature. Alternatively, by applying a saturating external magnetic field, ferromagnetic peaks can be made to appear or vanish, since magnetic scattering amplitudes vanish if moments align along Q, to cleanly separate atomic and magnetic structure information.

Figure 12: 
Left: Structures and neutron powder diffraction data of hexagonal MnTe and 5%-Li:MnTe, with basal and axial spins, respectively. The [0001] magnetic peak is present only when moments are perpendicular to the c-axis. Adapted from ref 130 by kind permission of the American Physical Society. Right: Diffuse scattering from Ho2Ti2O7 spin ice measured with polarization analysis in the spin-flip channel (A), non-spin-flip channel (B), and their sum (C) as would be observed without polarization analysis. Adapted from ref 131 by kind permission of AAAS.
Figure 12:

Left: Structures and neutron powder diffraction data of hexagonal MnTe and 5%-Li:MnTe, with basal and axial spins, respectively. The [0001] magnetic peak is present only when moments are perpendicular to the c-axis. Adapted from ref 130 by kind permission of the American Physical Society. Right: Diffuse scattering from Ho2Ti2O7 spin ice measured with polarization analysis in the spin-flip channel (A), non-spin-flip channel (B), and their sum (C) as would be observed without polarization analysis. Adapted from ref 131 by kind permission of AAAS.

Having no new periodicity in a magnetic structure does not necessarily imply ferromagnetism. So-called k = 0 structures, for which the unit cells of lattice and spin structure coincide, can also be antiferromagnetic. A most useful convention is to describe magnetic structures by their propagation vectors k that capture the magnetic periodicity. The definition of a multi-k structure requires more than a single k vector. For high-symmetry lattices it can be difficult to distinguish a multi-k structure from equal domain populations of a single-k structure. These structures have gathered much attention because they yield real-space spin textures, i.e., skyrmions. Unambiguous identification of a multi-k magnetic structure requires additional information either from local-probe methods 132 or from the spin-dynamics. 133 A rational or irrational ratio of k to the reciprocal lattice parameters corresponds to a commensurate and incommensurate magnetic structure, respectively. For structural analysis, these can be treated similarly and are typically related to spiral structures. Extending our experimental repertoire to add polarization analysis (see Section 3.3) may facilitate the separation of magnetic from nuclear scattering and help to distinguish anisotropies in spin correlations. In view of the more complex structural patterns of spin spiral structures, neutron polarization analysis of diffraction from single crystals clearly opens new possibilities, revealing the vector chirality, M (−k) × M (k), i.e., the turning of helices and cycloids that remains hidden to unpolarized neutrons. The most precise tool is spherical polarimetry, see the recently developed software Mag2Pol. 134

Deviations away from the average magnetic structure yield diffuse magnetic scattering. An exemplary origin of such diffuse scattering is the magnetic disorder arising from geometric frustration of magnetic interactions. Antiferromagnetically coupled moments on an atomic triangle are an important example because triangles are a most common structural motif and building block in nature. Local spin anisotropies can impose strong frustration also on ferromagnetic, dipolar interactions. A most prominent example is the diffuse magnetic scattering from a single crystal of the ternary oxide Ho2Ti2O7, 131 also going under the name “spin-ice”. This magnetic analogue of water ice shows the same topology of disorder and ground state entropy. 135 In this spin structure, violations of the so-called ice rules 136 create remarkable defects, magnetic monopoles, at least in a topological sense. The distinct features are pinch points arising from the effective long-range Coulomb-like interactions, appearing as ridges of magnetic diffuse scattering through the (002) reciprocal lattice points. These are revealed using polarization analysis, with polarization perpendicular to the scattering plane to measure in-plane moments in spin-flip mode, see Figure 12, right.

3.5 In situ and operando studies, sample environment

Many characterization techniques yield a rather static picture of matter at an atomic level. While this does correspond to highly valuable information, structures will nonetheless change over time, i.e., dynamics, phase transitions and chemical reactions are not covered. In situ measurements collecting data while external parameters (e.g., T, p, gas flow, electric or magnetic field) change can shed light on such processes. If the function or wear of devices under realistic operating conditions are investigated in a time-resolved manner, we also speak of operando studies. They give a realistic picture of materials and their changes at an atomistic level over time in devices under “real-world” conditions. This relates to, for instance, mapping hydrogen uptake in hydrogen storage materials, lithium intercalation in battery materials, oxygen defects in oxygen ion conductors for fuel cells, changes in microstructure in alloys, or phase formation in chemical synthesis, just to name a few examples. 137 , 138 Such deep insights into time-resolved processes through operando studies contribute significantly to the optimization of whatever devices.

Obviously, in situ and operando investigations rely on controlling the external parameters to be varied. This is provided by sample environment, which may obstruct the primary or diffracted beam. Since neutrons interact weakly with most materials and have a high penetration depth, they allow for bulky sample environment and are thus ideally suited for in situ and operando studies. In addition, neutron scattering is non-destructive but highly sensitive for light elements, especially hydrogen, lithium, nitrogen and oxygen, and so neutrons may probe many interesting materials properties such as crystal structure, magnetism, atomic diffusion or vibrational properties. As light atoms such as H or Li are highly relevant for “energy materials”, neutrons are important in this respect, too.

Eventually, the high transparency of most sample environments for neutrons also facilitates the use of cryostats down to the milli-Kelvin range, not achievable by synchrotron radiation.

The sample environment for controlling the external parameters may be of a general nature and available at many neutron beamlines, e.g., furnaces (up to 2,000 K or more), cryostats (T down to mK 139 ), magnets (up to 26 T), pressure cells (up to 100 GPa 140 , 141 ), or highly specialized equipment such as tensile stress testing machines, electrochemical cells, 142 or gas-pressure cells. 143 The combination of two or more parameters to be varied is also quite common, such as cryomagnet, pressure cell in a cryostat, electric fields in a furnace or heating in gas flow. To give even deeper insight, in situ neutron diffraction may be combined with simultaneous characterization, for example by spectroscopy (Raman, IR, mass spectrometry).

Let us look at a few examples to demonstrate the power of in situ and operando neutron diffraction. Rechargeable batteries are often based on lithium (de)intercalation into solid electrode materials, for which neutrons are an excellent probe. In situ and operando investigations were crucial in explaining failure mechanisms in lithium manganite-based rechargeable lithium-ion batteries. An unfavorable capacity fading observed for LiMn2O4, but not for lithium-rich Li1+x Mn2–x O4, was tracked down to different reaction pathways. Samples with x = 0.0 suffer large volume changes upon formation of three distinct phases, while samples with x = 0.1 show a continuous lithium (de)intercalation within the same phase (Figure 13, left) resulting in a lower mechanical stress during cycling and, therefore, much less pronounced capacity fading for the non-stoichiometric Li1+x Mn2–x O4. 144 Due to the large penetration depth, such operando investigations may be done on commercial battery packs without any disassembly or further preparation, thus providing real-world conditions, 142 as said before. In hydrogen storage materials, additives are often used in order to improve reaction rates of hydrogen uptake and release. There are numerous examples where in situ neutron diffraction could explain these findings by identifying reaction intermediates, e.g., AlB2 for Al additives to LiBH4, or mixed imides Li2Mg(ND)2 and Li2Mg2(ND)3 in the Li–N–H system. Hydrogen incorporation into magnetic materials is another example, where in situ neutron experiments as a function of temperature and hydrogen gas pressure can reveal the complex interplay between hydrogen content, atomic and magnetic order (Figure 13, right). 145 Nitrides are a natural target for neutron diffraction due to the large scattering contribution of nitrogen atoms for thermal neutrons. For example, following the ammonolysis of iron powder by time-resolved in situ neutron diffraction unraveled the formation mechanism of the giant magnetic moment material α’’-Fe16N2. In combination with thermal analysis, forming the iron-rich nitride was found to be kinetically controlled. This helped to optimize the synthesis and increase the yield of this metastable solid to above 90 %. 137

Figure 13: 
Left: in situ neutron powder diffraction of Li1+x
Mn2–x
O4 cathode materials upon lithium deintercalation in a lithium-ion battery revealing three distinct phases for x = 0 and the continuous lithium deintercalation for x = 0.1, thus explaining lower mechanical stress and less capacity fading in the latter. Adapted from ref 144 by kind permission of the American Chemical Society. Right: in situ neutron powder diffraction data (λ = 1.8680(2) Å) of the reaction of FePd3 with deuterium gas as a function of temperature and deuterium gas pressure yield deuterium content, atomic and magnetic order by Rietveld analysis, revealing their complex interplay in this ferromagnet. Adapted from ref 145 by CC BY.
Figure 13:

Left: in situ neutron powder diffraction of Li1+x Mn2–x O4 cathode materials upon lithium deintercalation in a lithium-ion battery revealing three distinct phases for x = 0 and the continuous lithium deintercalation for x = 0.1, thus explaining lower mechanical stress and less capacity fading in the latter. Adapted from ref 144 by kind permission of the American Chemical Society. Right: in situ neutron powder diffraction data (λ = 1.8680(2) Å) of the reaction of FePd3 with deuterium gas as a function of temperature and deuterium gas pressure yield deuterium content, atomic and magnetic order by Rietveld analysis, revealing their complex interplay in this ferromagnet. Adapted from ref 145 by CC BY.

Many more examples could be described in detail, for example exotic behavior of matter under extreme conditions, mimicking conditions found deep in the Earth’s mantle in geochemistry, studying magnetic ordering at low temperature or the kinetics of chemical reactions.

The time resolution of the method used has to match that of the reaction to be followed. For in situ and operando neutron diffraction the resolution typically ranges from milliseconds to hours and is largely determined by the neutron flux on the sample and the detector efficiency. This obviously limits studies with high time resolution to specialized high-flux instruments. For fast reactions (within seconds) usually only qualitative phase analysis and lattice parameter determination may be performed, while for slower ones (within minutes) the full Rietveld analysis yielding complete crystal structures is feasible given modern neutron diffraction instruments. 138 Future developments in neutron instrumentation will allow even better time resolution. Due to the lower brilliance of the source, present-day neutron diffraction cannot compete with synchrotron diffraction in terms of time resolution in general, but often provides extremely valuable information in the accessible parameter space.

Hence, in situ and operando neutron diffraction are powerful tools for investigating the structural properties and behavior of materials. These techniques yield detailed information on phase transitions and chemical processes including phase formation, reaction intermediates and their crystal and micro-structures in real time, providing in-depth insights into their physical and chemical properties. This not only applies to basic research but also to more applied questions such as industrial processes or the function and wear of functional materials and devices.

3.6 High pressure

High-pressure experiments are an attractive tool in a rather broad research area, targeting all kinds of physical probes. Unlike chemical pressure – created by a too small atom sitting on a too large site, hence volume contraction –, high pressure is a clean and controllable external parameter. Squeezing the interatomic distance can provide a means to control and understand various physical properties, such as superconducting transitions, and even start pressure-dependent chemical reactions. 146 It is also an indispensable tool in the search for novel materials and planetary science. Basically, the high-pressure experiments are incompatible with neutron experiments requiring larger sample volumes, since the pressure generated is determined by the force applied divided by the area, which inevitably reduces the volume of the sample. Developing unique high-pressure cells, however, has facilitated its application to neutron experiments.

In neutron diffraction under high pressure, the following three points are essential. First, the sample volume should be kept as large as possible to counteract the weak interaction of neutrons with matter. Second, detector coverage should be secured to acquire the signal from the small sample. Finally, background countermeasures are necessary to reduce parasitic scattering from cells, anvils, and pressure-transmitting media around the sample.

Let us take the example of a powder diffraction experiment using a Paris–Edinburgh (PE) cell, 140 , 147 which has expanded the possibilities of neutron experiments under high pressure (Figure 14). First, using the anvils with an indentation at the tip, a sufficient sample volume of 0.088 mL is ensured for neutron experiments up to 10 GPa. Next, the detector coverage is secured by a frame design with limited shadow area and the anvils that ensure a ±7° vertical window even under high loads. Finally, measures to record clean patterns are employed according to the source used and the purpose of the experiment. In a reactor source, the angle-dispersive method is applied with a “through the gasket” geometry. The use of cubic boron nitride, BN, an anvil material that does not show strong Bragg peaks, contributes to a clean diffraction pattern. On the other hand, the TOF diffraction at a pulsed neutron source can access a wide energy range even with a limited diffraction angle. In this case, the “through the anvil” geometry is generally applied. The prospective area contributing to the signal is confined by the incident collimator and the anvil gap, effectively reducing the parasitic signal from surrounding materials. The PE cells have been installed in neutron facilities worldwide and allow us to conduct Rietveld refinement as well as PDF analysis under extreme conditions.

Figure 14: 
Paris-Edinburgh cell and scattering geometries with the inset showing a magnified view of the sample and anvils.
Figure 14:

Paris-Edinburgh cell and scattering geometries with the inset showing a magnified view of the sample and anvils.

In addition to the PE cell, various high-pressure devices have been developed. For example, a large single crystal diamond anvil cell (DAC) was developed for SNAP at the SNS, and the pressure generated is exceeding 100 GPa. 141 , 148 The DAC enables the gas to be sealed inside the sample chamber, so it is expected to play a significant role in the future structural elucidation of superconducting hydrides in which hydrogen plays an important role. 149 At PLANET at J-PARC, a large volume multi-anvil press for high-temperature and high-pressure experiments has been installed, 150 and research on iron hydrides is conducted to estimate the amount of hydrogen contained in the Earth’s core. 151 A very chemical example is given by the high-pressure (2 GPa) synthesis and structural elucidation of the ubiquitous “non-existing” compound carbonic acid, H2CO3, for which a deuterated sample was used for diffraction in a specially built 0.4 mL hybrid clamped cell, followed by density functional-based structure solution. 152 While most high-pressure studies are on powder samples, some examples of single-crystal structural studies should be mentioned: a single-crystal DAC that can be used complementary for both X-rays and neutrons, 153 and a DAC with polycrystalline diamonds as an anvil to overcome the problem of cell absorption correction. 154 In addition, neutron polarimetry experiments at ILL succeeded in determining a complex spiral magnetic structure in a multiferroic material at 4 GPa using a nonmagnetic high-pressure cell. 155 With the development of the unique cell, high-pressure neutron diffraction experiments are expected to continue to expand their range of applications.

4 Challenges and outlook

It will not surprise the reader too much that the authors of this primer are very enthusiastic about the superior attributes of neutron diffraction. The inherent challenges of the method, however, which were already covered in the individual sections before, should be mentioned one more time. First of all, the strength of the physical interaction of neutrons with matter turns out to be rather small such that plenty of neutrons and/or large sample amounts are needed – but thanks to angle-independent scattering lengths, there is no inherent drop of scattered intensity for increasing angles in contrast to X-rays. Second, the neutrons must be extracted from the atomic nuclei, and that requires large-scale research facilities and, for that very reason, big chunks of money, on the order of billions (euros, dollars, pounds, francs), considerably more than needed for the handy X-ray diffractometer in the basement of any good university institute. This sad “characteristic” of costliness is shared by neutron research with synchrotron research, but only for neutrons one needs to deal with “atoms” or, even worse, “nuclei”, and any such research alluding to “atomic” things is viewed skeptically or even ideologically here and there. Because of that, the political reputation of nuclear reactors for both energy production and research has suffered severely in the last decades, with important consequences: lack of neutrons. Although spallation sources evolved to substitute or complement those research reactors in some countries worldwide, the amount and speed of replacement has not been sufficient recently to keep up with the needs of a vital neutron research community.

But where there is danger, the saving also grows (Hölderlin). We have already mentioned that releasing neutrons does not require a nuclear reactor, but can be done more efficiently with accelerator-driven pulsed-mode sources; for example, one may compare the competitive efficiency of the instruments at the Second target Station at ISIS operating at 50 kW with the 50 MW ILL reactor. There is a new focus on producing neutrons by means of low-energy nuclear reactions, which sets the footing for the future construction of local neutron machines. At the time of writing this primer, various countries are developing such Compact Accelerator-driven Neutron Sources (CANS), not only serving as local facilities but also of fundamental importance when considering educating and training future neutron researchers or developing sample environments. Additionally, so called High Current Accelerator-driven Neutron Sources (HiCANS) generate neutrons in a pulsed mode with peak brilliances competitive to instruments at nuclear reactors. HiCANS can readily serve as user facilities and are officially treated as accelerator sources – just like synchrotrons – thus they do not fall under nuclear regulations, triggering substantial hope for a revival of a sustainable neutron landscape in the near future. At the same time, the authors look forward, with great expectations, to the completion of the world’s most modern and powerful spallation source: the ESS, located in Lund (Sweden), will certainly attract many neutron researchers from everywhere, and at the same time one may anticipate many new discoveries, which are only possible through a huge flux and, not to forget, greater brilliance.

Admittedly, sometimes neutron diffraction is presented as a kind of “stopgap”, necessary only when X-ray diffraction fails or is exhausted, i.e., because of too light nuclei, too similar nuclei, magnetic phenomena, etc. We, too, have mentioned such issues, after all they are simply true. But this way of thinking falls short, especially since the neutron alone brings a combination of properties (wave-like, mass-like, and a miniature magnet) that no other probe shares. In this respect, neutron diffraction is virtually unique and, when available, can replace a whole range of other techniques, completely. For diffraction, the wave-like nature is crucial, and for the elucidation of magnetic structures, the magnetic moment of the neutron comes in handy. Without magnetic neutron diffraction, we would know practically nothing about magnetic structures in terms of drawing a simple picture with up/down arrows for magnetic moments, and no hard disk (and the information society that depends on it) could ever have been built. This is an important reminder for stakeholders that decide on the construction of future research centers. In this connection we also want to mention an aspect which was neglected before, on purpose. Precisely because the neutron has a mass, it can exchange a little momentum when it comes into contact with matter (the crystal receives a non-elastic kick and also exchanges energy), so in this respect neutron research penetrates into the field of spectroscopy. In principle, a diffractometer can be converted into a spectrometer, thus comparable to an IR or Raman instrument, but this time based on neutrons, and, beyond revealing where atoms are, seeing how they move. The neutron can literally do everything.

Coming back to the new generations of neutron sources, the foreseeable prospects of neutron research are downright excellent. The infinite advantage of neutron research (isotope specific, magnetic, momentum and energy transfer and, hence, spectroscopy) will be central to answer a plethora of important scientific questions, for instance to the structure and dynamics of light atoms in energy conversion and storage materials, magnetic matter, or even protein and virus structures. So, one can expect enormous gain of knowledge through “ordinary” powder diffraction, in single-crystal diffraction, and also in texture measurements. We have stressed the self-evident relationship between spallation sources and the TOF neutron-diffraction technique: at the same time, it must be ensured that the data processing itself keeps pace with the enormous increase in data volumes. The decisive algorithms for this endeavor (among others: two- and multidimensional Rietveld) are already ready. Data treatment routines based on artificial intelligence (AI)/machine learning (ML) techniques are in development, too, but would need a substantial funding with respect to neutron scattering and instrumentation research to make use of the general boost in the AI/ML fields.

Furthermore, the higher Q-domain will bring a huge information gain for nanomaterials, diffuse scattering as well as magnetic studies in general. Whereas in the past studies with polarized neutrons suffered from an additional decrease in counting rates, this problem will be considerably less critical at the new sources. The same is true for scientific research using in situ and operando conditions, and equally so for high-pressure experiments. If measurements are to be made quickly (in situ, operando) or with small quantities (high pressure), high neutron counting rates are of crucial importance, and they will be available. These fantastic prospects in future neutron diffraction do not absolve the individual scientist from making the best possible or “right” design for the experiment (this demands creativity in the first place), but it does make neutron diffraction, regardless of the unrivaled properties of the neutron, an extremely desirable technique. If it did not exist, it would have to be invented.

The authors would be happy if this primer could get young people excited about neutron diffraction, especially since it is one of the most powerful techniques available, and it is gaining momentum, again. Please start planning your first experiment today, preferably by contacting a major research facility (Tables 2 and 3) and the local scientist at that site. While access to beamtime is competitive and merit-based, it is completely free. Good luck!


Corresponding author: Richard Dronskowski, Institute of Inorganic Chemistry, RWTH Aachen University, 52056 Aachen, Germany, E-mail:

Acknowledgment

RD and AH gratefully acknowledge the continuous support by the German Federal Ministry of Research and Education (BMBF), also in terms of open-access funding via ErUM-Pro project 05K22PA2. The work at Oak Ridge National Laboratory was supported by the Department of Energy, Office of Basic Energy Science, Materials Sciences and Engineering Division.

  1. Research ethics: Not applicable.

  2. Author contribution: All the authors have contributed equally.

  3. Competing interests: Not applicable.

  4. Research funding: Federal Ministry of Research and Education (BMBF), ErUM-Pro project 05K22PA2, Germany; Department of Energy, Office of Basic Energy Science, Materials Sciences and Engineering Division, USA.

  5. Data availability: Not applicable.

References

1. Carpenter, J. M.; Loong, C.-K. Elements of Slow-Neutron Scattering; Cambridge University Press: Cambridge, UK, 2015.10.1017/CBO9781139029315Suche in Google Scholar

2. Brückel, T.; Gutberlet, T.; Zakalek, P. Brillante Neutronenstrahlen – Eine neue Generation von Neutronenquellen für Wissenschaft und Industrie. Phys. J. 2023, 22, 7.Suche in Google Scholar

3. Zakalek, P.; Doege, P.-E.; Baggemann, J.; Mauerhofer, E.; Brückel, T. Energy and Target Material Dependence of the Neutron Yield Induced by Proton and Deuteron Bombardment. EPJ Web Conf. 2020, 231, 03006; https://doi.org/10.1051/epjconf/202023103006.Suche in Google Scholar

4. Brückel, T.; Gutberlet, T.; Schmidt, S.; Alba-Simionesco, C.; Ott, F.; Menelle, A. Low Energy Accelerator-Driven Neutron Facilities – A Prospect for a Brighter Future for Research with Neutrons. Neutron News 2020, 31, 2–4; https://doi.org/10.1080/10448632.2020.1819125.Suche in Google Scholar

5. Squires, G. L. Introduction to the Theory of Thermal Neutron Scattering; Cambridge University Press: Cambridge, UK, 2012.10.1017/CBO9781139107808Suche in Google Scholar

6. Dianoux, A.-J.; Lander, G. Neutron Data Booklet; Institut Laue-Langevin: Grenoble, 2003.Suche in Google Scholar

7. Brown, P. J. Magnetic form factors. In International Tables for Crystallography. Vol. C, Section 4.4.5; Prince, E., Ed.; Kluwer Academic Publishers: Dordrecht, 2004; pp 454–461.Suche in Google Scholar

8. Kohlmann, H. Solid-State Structures and Properties of Europium and Samarium Hydrides. Eur. J. Inorg. Chem. 2010, 2582–2593; https://doi.org/10.1002/ejic.201000107.Suche in Google Scholar

9. Ryan, D. H.; Cranswick, L. M. D. Flat-Plate Single-Crystal Silicon Sample Holders for Neutron Powder Diffraction Studies of Highly Absorbing Gadolinium Compounds. J. Appl. Crystallogr. 2008, 41, 198–205; https://doi.org/10.1107/s0021889807065806.Suche in Google Scholar

10. Jacobs, P.; Houben, A.; Schweika, W.; Tchougréeff, A. L.; Dronskowski, R. A Rietveld Refinement Method for Angular- and Wavelength-Dispersive Neutron Time-of-Flight Powder-Diffraction Data. J. Appl. Crystallogr. 2015, 48, 1627–1636; https://doi.org/10.1107/s1600576715016520.Suche in Google Scholar

11. Young, R. A.; Ed.; The Rietveld Method; Oxford University Press: Oxford, UK, 1993.10.1093/oso/9780198555773.001.0001Suche in Google Scholar

12. Chadwick, J. Possible Existence of a Neutron. Nature 1932, 129, 312; https://doi.org/10.1038/129312a0.Suche in Google Scholar

13. Elsasser, W. Sur la Diffraction des Neutrons Lents par les Substances Cristallines. C. R. Acad. Sci. 1936, 202, 1029.Suche in Google Scholar

14. Preiswerk, P.; von Halban, H. Preuve Experimental de la Diffraction des Neutrons. C. R. Acad. Sci. 1936, 203, 73–75.Suche in Google Scholar

15. Mitchell, D. P.; Powers, P. N. Bragg Reflection of Slow Neutrons. Phys. Rev. 1936, 50, 486–487; https://doi.org/10.1103/physrev.50.486.2.Suche in Google Scholar

16. Wilkinson, M. Early History of Neutron Scattering at Oak Ridge. Physica B+C 1986, 137, 3–16; https://doi.org/10.1016/0378-4363(86)90306-2.Suche in Google Scholar

17. Mason, T. E.; Gawne, T. J.; Nagler, S. E.; Nestor, M. B.; Carpenter, J. M. The Early Development of Neutron Diffraction: Science in the Wings of the Manhattan Project. Acta Crystallogr. A 2013, 69, 37–44; https://doi.org/10.1107/s0108767312036021.Suche in Google Scholar

18. Wollan, E. O.; Shull, C. The Diffraction of Neutrons by Crystalline Powders; US Atomic Energy Commission: Washington, DC, 1948.10.1103/PhysRev.73.830Suche in Google Scholar

19. Caglioti, G.; Paoletti, A.; Ricci, F. P. Choice of Collimators for a Crystal Spectrometer for Neutron Diffraction. Nucl. Instr. 1958, 3, 223–228; https://doi.org/10.1016/0369-643x(58)90029-x.Suche in Google Scholar

20. Hewat, A. W. Design for a Conventional High-Resolution Neutron Powder Diffractometer. Nucl. Instrum. Methods 1975, 127, 361–370; https://doi.org/10.1016/s0029-554x(75)80006-1.Suche in Google Scholar

21. Shull, C. G.; Wollan, E. O.; Morton, G. A.; Davidson, W. L. Neutron Diffraction Studies of NaH and NaD. Phys. Rev. 1948, 73, 842; https://doi.org/10.1103/physrev.73.842.Suche in Google Scholar

22. Rietveld, H. M. Line Profiles of Neutron Powder-Diffraction Peaks for Structure Refinement. Acta Crystallogr. 1967, 22, 151–152; https://doi.org/10.1107/s0365110x67000234.Suche in Google Scholar

23. Rietveld, H. M. A Profile Refinement Method for Nuclear and Magnetic Structures. J. Appl. Crystallogr. 1969, 2, 65; https://doi.org/10.1107/s0021889869006558.Suche in Google Scholar

24. Loopstra, B.; Rietveld, H. The Structure of Some Alkaline-Earth Metal Uranates. Acta Crystallogr. B 1969, 25, 787–791; https://doi.org/10.1107/s0567740869002974.Suche in Google Scholar

25. van Laar, B.; Schenk, H. The Development of Powder Profile Refinement at the Reactor Centre Netherlands at Petten. Acta Crystallogr. A 2018, 74, 88–92; https://doi.org/10.1107/s2053273317018435.Suche in Google Scholar

26. Kaduk, J. A.; Billinge, S. J. L.; Dinnebier, R. E.; Henderson, N.; Madsen, I.; Černý, R.; Leoni, M.; Lutterotti, L.; Thakral, S.; Chateigner, D. Powder Diffraction. Nat. Rev. Methods Prim. 2021, 1, 77; https://doi.org/10.1038/s43586-021-00074-7.Suche in Google Scholar

27. Soller, W. A New Precision X-Ray Spectrometer. Phys. Rev. 1924, 24, 158–167; https://doi.org/10.1103/physrev.24.158.Suche in Google Scholar

28. Bacon, G. E. Neutron Diffraction; Clarendon Press: Oxford, 1975.10.1038/257360a0Suche in Google Scholar

29. Kisi, E. H.; Howard, C. J. Applications of Neutron Powder Diffraction; Oxford University Press: Oxford, UK, 2012.Suche in Google Scholar

30. Howard, C. J.; Kisi, E. H. Neutron powder diffraction. In International Tables for Crystallography Volume H: Powder DiffractionSection 2.3; Gilmore, C. J., Kaduk, J. A., Schenk, H., Eds.; Wiley: New York, 2019; pp 66–101.10.1107/97809553602060000938Suche in Google Scholar

31. von Dreele, R. B. Neutron powder diffraction. In Modern Powder Diffraction; Bish, D. L., Post, J. E., Eds.; De Gruyter: Berlin, Germany, 1989.10.1515/9781501509018-014Suche in Google Scholar

32. Copley, J. R. D. NIST Recommended Practice Guide: The Fundamentals of Neutron Powder Diffraction; National Institute of Standards and Technology: Gaithersburg, MD, 2001.10.6028/NBS.SP.960-2Suche in Google Scholar

33. Hewat, A. D2B, A New High Resolution Neutron Powder Diffractormeter at ILL Grenboble. Mat. Sci. Forum 1986, 9, 69–80; https://doi.org/10.4028/www.scientific.net/msf.9.69.Suche in Google Scholar

34. Hansen, T. C.; Henry, P. F.; Fischer, H. E.; Torregrossa, J.; Convert, P. The D20 Instrument at the ILL: A Versatile High-Intensity Two-Axis Neutron Diffractometer. Meas. Sci. Tech. 2008, 19, 034001; https://doi.org/10.1088/0957-0233/19/3/034001.Suche in Google Scholar

35. Hoelzel, M.; Senyshyn, A.; Juenke, N.; Boysen, H.; Schmahl, W.; Fuess, H. High-Resolution Neutron Powder Diffractometer SPODI at Research Reactor FRM II. Nucl. Instrum. Methods Phys. Res. A 2012, 667, 32–37; https://doi.org/10.1016/j.nima.2011.11.070.Suche in Google Scholar

36. Heere, M.; Mühlbauer, M. J.; Schökel, A.; Knapp, M.; Ehrenberg, H.; Senyshyn, A. Energy Research with Neutrons (ErwiN) and Installation of a Fast Neutron Powder Diffraction Option at the MLZ, Germany. J. Appl. Crystallogr. 2018, 51, 591–595; https://doi.org/10.1107/s1600576718004223.Suche in Google Scholar

37. Fischer, P.; Frey, G.; Koch, M.; Könnecke, M.; Pomjakushin, V.; Schefer, J.; Thut, R.; Schlumpf, N.; Bürge, R.; Greuter, U.; Bondt, S.; Berruyer, E. High-Resolution Powder Diffractometer HRPT for Thermal Neutrons at SINQ. Phys. B Condens. Matter 2000, 276, 146–147; https://doi.org/10.1016/s0921-4526(99)01399-x.Suche in Google Scholar

38. van Eijck, L.; Cussen, L.; Sykora, G.; Schooneveld, E.; Rhodes, N.; Van Well, A.; Pappas, C. Design and Performance of a Novel Neutron Powder Diffractometer: PEARL at TU Delft. J. Appl. Crystallogr. 2016, 49, 1398–1401; https://doi.org/10.1107/s160057671601089x.Suche in Google Scholar

39. Frontzek, M. D.; Andrews, K. M.; Jones, A. B.; Chakoumakos, B. C.; Fernandez-Baca, J. A. The Wide Angle Neutron Diffractometer Squared (WAND2)-Possibilities and Future. Phys. B Condens. Matter 2018, 551, 464–467; https://doi.org/10.1016/j.physb.2017.12.027.Suche in Google Scholar

40. Cappelletti, R. L.; Glinka, C. J.; Krueger, S.; Lindstrom, R.; Lynn, J. W.; Prask, H. J.; Prince, E.; Rush, J. J.; Rowe, J. M.; Satija, S. K.; Toby, B.; Tsai, A.; Udovic, T. Materials Research with Neutrons at NIST. J. Res. Nat. Inst. Stand. Tech. 2001, 106, 187; https://doi.org/10.6028/jres.106.008.Suche in Google Scholar PubMed PubMed Central

41. Avdeev, M.; Hester, J. R. ECHIDNA: A Decade of High-Resolution Neutron Powder Diffraction at OPAL. J. Appl. Crystallogr. 2018, 51, 1597–1604; https://doi.org/10.1107/s1600576718014048.Suche in Google Scholar

42. Studer, A. J.; Hagen, M. E.; Noakes, T. J. Wombat: The High-Intensity Powder Diffractometer at the OPAL Reactor. Phys. B Condens. Matter 2006, 385, 1013–1015; https://doi.org/10.1016/j.physb.2006.05.323.Suche in Google Scholar

43. Ohoyama, K.; Kanouchi, T.; Nemoto, K.; Ohashi, M.; Kajitani, T.; Yamaguchi, Y. HERMES, The High-Efficiency Multipurpose Powder Diffractometer. Phys. B Condens. Matter 1997, 241, 213–215; https://doi.org/10.1016/s0921-4526(97)00554-1.Suche in Google Scholar

44. Li, T.; Wu, M.; Jiao, X.; Sun, K.; Chen, D. Current Status and Future Prospect of Neutron Facilities at China Advanced Research Reactor. Nucl. Phys. Rev. 2020, 37, 364–376.Suche in Google Scholar

45. Xia, Y.; Li, H.; Cao, X.; Chen, X.; Zhang, C.; Wang, Y.; Huang, C.; Wu, Z.; Xie, C.; Sun, D. Upgrade of Xuanwu: A Dual-Mode Neutron Powder Diffractometer at CMRR. Nucl. Instrum. Methods Phys. Res. A 2022, 1042, 167452; https://doi.org/10.1016/j.nima.2022.167452.Suche in Google Scholar

46. Xie, L.; Chen, X.; Fang, L.; Sun, G.; Xie, C.; Chen, B.; Li, H.; Ulyanov, V.; Solovei, V.; Kolkhidashvili, M. Fenghuang: High-Intensity Multi-Section Neutron Powder Diffractometer at CMRR. Nucl. Instrum. Methods Phys. Res. A 2019, 915, 31–35; https://doi.org/10.1016/j.nima.2018.10.002.Suche in Google Scholar

47. Shim, H.-S.; Lee, C.-H.; Seong, B.-S.; Lee, J.-S. The status of neutron beam utilization in Korea, JAERI-Cof 99-012. In Proc. of the 1998 Workshop on the Utilization of Research Reactors, 1999; pp 52–60.Suche in Google Scholar

48. Lee, C.-H.; Em, V.; Moon, M.-K.; Hong, K.-P.; Cheon, J.-K.; Choi, Y.-H.; Nam, U.-W.; Kong, K.-N. High-Intensity Multi-psd Powder Diffractometer at the HANARO. Phys. B Condens. Matter 2006, 385, 1016–1018; https://doi.org/10.1016/j.physb.2006.05.324.Suche in Google Scholar

49. Krishna, P.; Chitra, R.; Choudhury, R.; Mishra, S. Structural Investigations of Crystalline and Amorphous Systems at Dhruva. Neutron News 2014, 25, 18–21; https://doi.org/10.1080/10448632.2014.870441.Suche in Google Scholar

50. Siruguri, V.; Babu, P.; Gupta, M.; Pimpale, A.; Goyal, P. A High Resolution Powder Diffractometer Using Focusing Optics. Pramana 2008, 71, 1197–1202; https://doi.org/10.1007/s12043-008-0246-2.Suche in Google Scholar

51. Chaplot, S.; Mukhopadhyay, R. Neutron Beam Research at Dhruva Reactor. Neutron News 2014, 25, 15–17; https://doi.org/10.1080/10448632.2014.870440.Suche in Google Scholar

52. Lowde, R. D. A New Rationale of Structure-Factor Measurement in Neutron-Diffraction Analysis. Acta Crystallogr. 1956, 9, 151–155; https://doi.org/10.1107/s0365110x56000346.Suche in Google Scholar

53. Buras, B. Note on the Integrated Intensity in the Time-of-Flight Method for Crystal Structure Investigations. Nukleonika 1963, 8, 259–260.Suche in Google Scholar

54. Buras, B.; Leciejewicz, J. A Time-of-Flight Method for Neutron Diffraction Crystal Structure Investigations. Nukleonika 1963, 8, 75–77.Suche in Google Scholar

55. Buras, B.; Leciejewicz, J.; Nitc, W.; Sosnowska, I.; Sosnowski, J.; Shapiro, F. The Time-of-Flight Method for Neutron Crystal Structure Investigations and its Possibilities in Connection With Very High Flux Reactors. Nukleonika 1964, 9, 523–535.Suche in Google Scholar

56. Aksenov, V. L.; Balagurov, A. M. Neutron Diffraction on Pulsed Sources. Phys. Usp. 2016, 59, 279–303; https://doi.org/10.3367/ufne.0186.201603e.0293.Suche in Google Scholar

57. Moore, M. J.; Kasper, J. S.; Menzel, J. H. Pulsed LINAC Neutron Diffraction. Nature 1968, 219, 848–849; https://doi.org/10.1038/219848a0.Suche in Google Scholar

58. Kimura, M.; Sugawara, M.; Oyamada, M.; Tsubota, T.; Tomiyoshi, S.; Watanabe, N.; Takeda, S.; Yamada, Y.; Hayasaka, H.; Watanabe, H.; Suzuki, T.; Ishikawa, Y.; Endoh, Y. Neutron Diffraction with an Electron Linac at Tohoku University; Improvement of Neutron Diffraction Using Electron Linac; Reactor Physics Experiments Using TOF (Time-of-Flight) Neutron Diffraction Method; Research Report of Laboratory of Nuclear Science (in Japanese); The Institute of Nuclear Science, Faculty of Science, Tohoku University: Tohoku, Vol. 1, 1968; pp 55–62; 81–104, 105–113.Suche in Google Scholar

59. Kimura, M.; Sugawara, M.; Oyamada, M.; Yamada, Y.; Tomiyoshi, S.; Suzuki, T.; Watanabe, N.; Takeda, S. Neutron Debye-Scherrer Diffraction Works Using a Linear Electron Accelerator. Nucl. Instrum. Methods 1969, 71, 102–110; https://doi.org/10.1016/0029-554x(69)90088-3.Suche in Google Scholar

60. Day, D. H.; Sinclair, R. N. Neutron Moderator Assemblies for Pulsed Thermal Neutron Time-of-Flight Experiments. Nucl. Instrum. Methods 1969, 72, 237–253; https://doi.org/10.1016/0029-554x(69)90509-6.Suche in Google Scholar

61. Steichele, E.; Arnold, P. A High Resolution Neutron Time of Flight Diffractometer. Phys. Lett. 1973, 44, 165–166; https://doi.org/10.1016/0375-9601(73)90866-9.Suche in Google Scholar

62. Huq, A.; Hodges, J. P.; Gourdon, O.; Heroux, L. POWGEN: A Third-Generation High-Resolution High-Throughput Powder Diffraction Instrument at the Spallation Neutron Source. Z. Kristallogr. Proc. 2011, 1, 127–135.10.1524/9783486991321-024Suche in Google Scholar

63. Jacobs, P.; Houben, A.; Tchougréeff, A. L.; Dronskowski, R. High-Resolution Neutron Diffraction Study of CuNCN: New Evidence of Structure Anomalies at Low Temperature. J. Chem. Phys. 2013, 139, 224707; https://doi.org/10.1063/1.4840555.Suche in Google Scholar PubMed

64. Tchougréeff, A. L.; Stoffel, R. P.; Houben, A.; Jacobs, P.; Dronskowski, R.; Pregelj, M.; Zorko, A.; Arcon, D.; Zaharko, O. Atomic Motions in the Layered Copper Pseudochalcogenide CuNCN Indicative of a Quantum Spin-Liquid Scenario. J. Phys.: Condens. Matter 2017, 29, 235701; https://doi.org/10.1088/1361-648x/aa6e73.Suche in Google Scholar PubMed

65. Rodríguez-Carvajal, J. Recent Advances in Magnetic Structure Determination by Neutron Powder Diffraction. Physica B 1993, 192, 55–69; https://doi.org/10.1016/0921-4526(93)90108-i.Suche in Google Scholar

66. von Dreele, B.; Toby, B. GSAS-II: The Genesis of a Modern Open-Source All Purpose Crystallography Software Package. J. Appl. Crystallogr. 2013, 46, 544–549; https://doi.org/10.1107/s0021889813003531.Suche in Google Scholar

67. Jacobs, P.; Houben, A.; Schweika, W.; Tchougréeff, A. L.; Dronskowski, R. Instrumental Resolution as a Function of Scattering Angle and Wavelength as Exemplified for the POWGEN Instrument. J. Appl. Crystallogr. 2017, 50, 866–875; https://doi.org/10.1107/s1600576717005398.Suche in Google Scholar PubMed PubMed Central

68. Modzel, G.; Henske, M.; Houben, A.; Klein, M.; Köhli, M.; Lennert, P.; Meven, M.; Schmidt, C. J.; Schmidt, U.; Schweika, W. Absolute Efficiency Measurements with the 10B Based Jalousie Detector. Nucl. Instrum. Methods A. 2014, 743, 90–95; https://doi.org/10.1016/j.nima.2014.01.007.Suche in Google Scholar

69. Houben, A.; Meinerzhagen, Y.; Nachtigall, N.; Jacobs, P.; Dronskowski, R. POWTEX Visits POWGEN. J. Appl. Crystallogr. 2023, 56, 633–642; https://doi.org/10.1107/s1600576723002819.Suche in Google Scholar PubMed PubMed Central

70. Meinerzhagen, Y.; Houben, A.; Dronskowski, R. PowderReduceP2D V1; docs.mantidproject.org/nightly/algorithms/PowderReduceP2D-v1.html; Algorithm Released with Mantid 6.2.0, 2021; https://doi.org/10.5286/SOFTWARE/MANTID6.2.Suche in Google Scholar

71. Arnold, O.; Bilheux, J. C.; Borreguero, J. M.; Buts, A.; Campbell, S. I.; Chapon, L.; Doucet, M.; Draper, N.; Ferraz Leal, R.; Gigg, M. A.; Lynch, V. E.; Markvardsen, A.; Mikkelson, D. J.; Mikkelson, R. L.; Miller, R.; Palmen, K.; Parker, P.; Passos, G.; Perring, T. G.; Peterson, P. F.; Ren, S.; Reuter, M. A.; Savici, A. T.; Taylor, J. W.; Taylor, R. J.; Tolchenov, R.; Zhou, W.; Zikovsky, J. Mantid ‒ Data Analysis and Visualization Package for Neutron Scattering and μSR Experiments. Nucl. Instrum. Methods A 2014, 764, 156–166; https://doi.org/10.1016/j.nima.2014.07.029.Suche in Google Scholar

72. Bunge, H. J. Advantages of Neutron Diffraction in Texture Analysis. Texture, Stress, Microstruct. 1989, 10, 265–307; https://doi.org/10.1155/tsm.10.265.Suche in Google Scholar

73. Brokmeier, H.-G.; Yi, S. B. Neutrons and Synchrotron Radiation in Engineering Materials: From Fundamentals to Applications; Wiley-VCH: Weinheim, 2017; pp 55–72.10.1002/9783527684489.ch3Suche in Google Scholar

74. Artioli, G. Crystallographic Texture Analysis of Archaeological Metals: Interpretation of Manufacturing Techniques. Appl. Phys. A 2007, 89, 899–908; https://doi.org/10.1007/s00339-007-4215-2.Suche in Google Scholar

75. Brokmeier, H.-G.; Gan, W. M.; Randau, C.; Völler, M.; Rebelo-Kornmeier, J.; Hofmann, M. Texture Analysis at Neutron Diffractometer STRESS-SPEC. Nucl. Instr. Methods A 2011, 642, 87–92; https://doi.org/10.1016/j.nima.2011.04.008.Suche in Google Scholar

76. Losko, A. S.; Vogel, S. C.; Reiche, H. M.; Nakotte, H. A Six-Axis Robotic Sample Changer for High-Throughput Neutron Powder Diffraction and Texture Measurements. J. Appl. Crystallogr. 2014, 47, 2109–2112; https://doi.org/10.1107/s1600576714021797.Suche in Google Scholar

77. Wenk, H. R. Neutron Diffraction Texture Analysis. Rev. Mineral. Geochem. 2006, 63, 399–426; https://doi.org/10.2138/rmg.2006.63.15.Suche in Google Scholar

78. Ivankina, T. I.; Matthies, S. On the Development of the Quantitative Texture Analysis and its Application in Solving Problems of the Earth Sciences. Phys. Part. Nucl. 2015, 46, 366–423; https://doi.org/10.1134/s1063779615030077.Suche in Google Scholar

79. Brokmeier, H. G. Advances and Applications of Neutron Texture Analysis, Texture, Stress, Microstruct. 1989, 33, 13–33; https://doi.org/10.1155/tsm.33.13.Suche in Google Scholar

80. Malmud, F.; Santisteban, J. R.; Vicente Alvarez, M. A.; Bolmaro, R.; Kellher, J.; Kabra, S.; Kockelmann, W. Texture Analysis with a Time-of-Flight Neutron Strain Scanner. J. Appl. Crystallogr. 2014, 47, 1337–1354; https://doi.org/10.1107/s1600576714012710.Suche in Google Scholar

81. Wenk, H. R.; Lutteroti, L.; Vogel, S. C. Rietveld Texture Analysis from TOF Neutron Diffraction Data. Powder Diffr. 2010, 25, 283–296; https://doi.org/10.1154/1.3479004.Suche in Google Scholar

82. Sawinski, P. K.; Meven, M.; Englert, U.; Dronskowski, R. Single-Crystal Neutron Diffraction Study on Guanidine, CN3H5. Cryst. Growth Des. 2013, 13, 1730–1735; https://doi.org/10.1021/cg400054k.Suche in Google Scholar

83. Jin, W. T.; Meven, M.; Sazonov, A. P.; Demirdis, S.; Su, Y.; Xiao, Y.; Bukowski, Z.; Nandi, S.; Brückel, T. Spin Reorientation of the Fe Moments in Eu0.5Ca0.5Fe2As2: Evidence for Strong Interplay of Eu and Fe Magnetism. Phys. Rev. B 2019, 99, 140402(R); https://doi.org/10.1103/physrevb.99.140402.Suche in Google Scholar

84. Schweika, W. XYZ Polarization Analysis of Diffuse Scattering from Single Crystals. J. Phys.: Confer. Ser. 2010, 211, 012026; https://doi.org/10.1088/1742-6596/211/1/012026.Suche in Google Scholar

85. Hutanu, V.; Meven, M.; Masalovich, S.; Heger, G.; Roth, G. 3He Spin Filters for Spherical Neutron Polarimetry at the Hot Neutrons Single Crystal Diffractometer POLI-HEiDi. J. Phys.: Confer. Ser. 2011, 294, 012012; https://doi.org/10.1088/1742-6596/294/1/012012.Suche in Google Scholar

86. Bärnighausen, H. Group-subgroup Relations Between Space Groups: A Useful Tool in Crystal Chemistry. Commun. Math. Chem. 1980, 9, 139–175.Suche in Google Scholar

87. Müller, U. Symmetry Relationships Between Crystal Structures; Oxford University Press: Oxford, UK, 2013.10.1093/acprof:oso/9780199669950.001.0001Suche in Google Scholar

88. Parsons, S. Introduction to Twinning. Acta Crystallogr. D 2003, 59, 1995–2003; https://doi.org/10.1107/s0907444903017657.Suche in Google Scholar PubMed

89. Redhammer, G. J.; Meven, M.; Ganschow, S.; Tippelt, G.; Rettenwander, D. Single-Crystal Neutron and X-Ray Diffraction Study of Garnet-Type Solid-State Electrolyte Li6La3ZrTaO12: An In Situ Temperature-Dependence Investigation (2.5 ≤ T ≤ 873 K). Acta Crystallogr. B 2021, 77, 123–130; https://doi.org/10.1107/s2052520620016145.Suche in Google Scholar

90. Gatta, G. D.; Vignola, P.; Meven, M.; Rinaldi, R. Neutron Diffraction in Gemology: Single-Crystal Diffraction Study of Brazilianite, NaAl3(PO4)2(OH)4. Amer. Mineral. 2013, 98, 1624–1630; https://doi.org/10.2138/am.2013.4476.Suche in Google Scholar

91. Gatta, G. D.; Hradil, K.; Meven, M. Where is the Hydrogen? In Elements, Special Issue on Exploring Earth and Planetary Materials with Neutrons (Cole, D. R. & Ross, N. L., Guest Editors). Elements 2021, 17, 163–168.10.2138/gselements.17.3.163Suche in Google Scholar

92. Mattauch, S. Untersuchung der strukturellen Phasenübergänge und Domänenbildung in den ferroischen Modellsubstanzen RbH2PO4 und RbD2PO4. Dissertation, RWTH Aachen University, 2002.Suche in Google Scholar

93. Truong, K.-N.; Merkens, C.; Meven, M.; Faßbänder, B.; Dronskowski, R.; Englert, U. Phase Transition and Proton Ordering at 50 K in 3-(pyridin-4-yl)pentane-2,4-dione. Acta Crystallogr. B 2017, 73, 1172–1178; https://doi.org/10.1107/s2052520617015591.Suche in Google Scholar

94. Niewa, R.; Czulucki, A.; Schmidt, M.; Auffermann, G.; Cichorek, T.; Meven, M.; Pedersen, B.; Steglich, F.; Kniep, R. Crystal Structure Investigations of ZrAsxSey (x > y, x+y ≤ 2) by Single Crystal Neutron Diffraction at 300 K, 25 K and 2.3 K. J. Solid State Chem. 2010, 183, 1309–1313; https://doi.org/10.1016/j.jssc.2010.04.006.Suche in Google Scholar

95. Gatta, G. D.; Redhammer, G. J.; Vignola, P.; Meven, M.; McIntyre, G. J. Single-crystal Neutron Diffraction and Mössbauer Spectroscopic Study of Hureaulite, (Mn,Fe)5(PO4)2(HPO4)2(H2O)4. Eur. J. Min. 2016, 28, 93–103; https://doi.org/10.1127/ejm/2015/0027-2464.Suche in Google Scholar

96. Welberry, R.; Whitfield, R. Single Crystal Diffuse Neutron Scattering. Quantum Beam Sci. 2018, 2, 2; https://doi.org/10.3390/qubs2010002.Suche in Google Scholar

97. Welberry, T. R. One Hundred Years of Diffuse X-Ray Scattering. Metall. Mater. Trans. A 2014, 45, 75–84; https://doi.org/10.1007/s11661-013-1889-2.Suche in Google Scholar

98. von Laue, M. Röntgenstrahlinterferenz und Mischkristalle. Ann. Phys. 1918, 361, 497–506; https://doi.org/10.1002/andp.19183611502.Suche in Google Scholar

99. Schweika, W. Disordered Alloys, Diffuse Scattering and Monte Carlo Simulations. Springer Tracts Mod. Phys., Vol. 141; Springer Verlag, Berlin, Heidelberg, New York, 1998.Suche in Google Scholar

100. Proffen, T.; Neder, R. B. DISCUS, A Program for Diffuse Scattering and Defect Structure Simulations. J. Appl. Crystallogr. 1997, 30, 171; https://doi.org/10.1107/s002188989600934x.Suche in Google Scholar

101. McGreevy, R. L.; Pusztai, L. Reverse Monte Carlo Simulation: A New Technique for the Determination of Disordered Structures. Mol. Sim. 1988, 1, 359–367; https://doi.org/10.1080/08927028808080958.Suche in Google Scholar

102. Keen, D. A.; Goodwin, A. L. The Crystallography of Correlated Disorder. Nature 2015, 521, 303–309; https://doi.org/10.1038/nature14453.Suche in Google Scholar PubMed

103. Paddison, J. A. M.; Stewart, J. R.; Goodwin, A. L. Spinvert: A Program for Refinement of Paramagnetic Diffuse Scattering Data. J. Phys.: Condens. Matter 2013, 25, 454220; https://doi.org/10.1088/0953-8984/25/45/454220.Suche in Google Scholar PubMed

104. Keen, D. A.; Gutmann, M. J.; Wilson, C. C. SXD – The Single-Crystal Diffractometer at the ISIS Spallation Neutron Source. J. Appl. Crystallogr. 2006, 39, 714–722; https://doi.org/10.1107/s0021889806025921.Suche in Google Scholar

105. Rosenkranz, S.; Osborn, R. Corelli: Efficient Single Crystal Diffraction with Elastic Discrimination. PRAMANA J. Phys. 2008, 71, 705–711; https://doi.org/10.1007/s12043-008-0259-x.Suche in Google Scholar

106. Egami, T.; Billinge, S. J. L. Underneath the Bragg Peaks. 2nd ed.; Pergamon: Oxford, UK, 2012.Suche in Google Scholar

107. Narten, A. H.; Thiessen, W. E.; Blum, L. Atom Pair Distribution Functions of Liquid Water at 25 °C from Neutron Diffraction. Science 1982, 217, 1033–1034; https://doi.org/10.1126/science.217.4564.1033.Suche in Google Scholar PubMed

108. Soper, A. K.; Neilson, G. W.; Enderby, J. E.; Howe, R. A. A Neutron Diffraction Study of Hydration Effects in Aqueous Solutions. J. Phys. C: Solid State Phys. 1977, 10, 1793; https://doi.org/10.1088/0022-3719/10/11/014.Suche in Google Scholar

109. Wright, A. C.; Clare, A. G.; Grimley, D. I.; Sinclair, R. N. Neutron Scattering Studies of Network Glasses. J. Non-Cryst. Solids 1989, 112, 33–47; https://doi.org/10.1016/0022-3093(89)90491-2.Suche in Google Scholar

110. Billinge, S. J. L.; Kanatzidis, M. G. Beyond Crystallography: The Study of Disorder, Nanocrystallinity and Crystallographically Challenged Materials with Pair Distribution Functions. Chem. Commun. 2004, 749–760; https://doi.org/10.1002/chin.200425250.Suche in Google Scholar

111. Dove, M. T.; Li, G. Review: Pair Distribution Functions from Neutron Total Scattering for the Study of Local Structure in Disordered Materials. Nuclear Anal. 2022, 1, 100037; https://doi.org/10.1016/j.nucana.2022.100037.Suche in Google Scholar

112. Luo, S.; Li, M.; Fung, V.; Sumpter, B. G.; Liu, J.; Wu, Z.; Page, K. New Insights into the Bulk and Surface Defect Structures of Ceria Nanocrystals from Neutron Scattering Study. Chem. Mater. 2021, 33, 3959–4397; https://doi.org/10.1021/acs.chemmater.1c00156.Suche in Google Scholar

113. Plekhanov, M. S.; Thomä, S. J. L.; Zobel, M.; Cuello, G. J.; Fischer, H. E.; Raskovalov, A. A.; Kuzmin, A. V. Correlating Proton Diffusion in Perovskite Triple-Conducting Oxides with Local and Defect Structure. Chem. Mater. 2022, 34, 4785–4794; https://doi.org/10.1021/acs.chemmater.2c01159.Suche in Google Scholar

114. Frandsen, B. A.; Yang, X.; Billinge, S. J. L. Magnetic Pair Distribution Function Analysis of Local Magnetic Correlations. Acta Crystallogr. A 2014, 70, 3–11; https://doi.org/10.1107/s2053273313033081.Suche in Google Scholar PubMed

115. Frandsen, B. A.; Billinge, S. J. L. Magnetic Structure Determination from the Magnetic Pair Distribution Function (mPDF): Ground State of MnO. Acta Crystallogr. A 2015, 71, 325–334; https://doi.org/10.1107/s205327331500306x.Suche in Google Scholar PubMed

116. Roth, N.; May, A. F.; Ye, F.; Chakoumakos, B. C.; Iversen, B. B. Model-Free Reconstruction of Magnetic Correlations in Frustrated Magnets. IUCrJ 2018, 5, 410–416; https://doi.org/10.1107/s2052252518006590.Suche in Google Scholar

117. Baral, R.; Christensen, J. A.; Hamilton, P. K.; Ye, F.; Chesnel, K.; Sparks, T. D.; Ward, R.; Yan, J.; McGuire, M. A.; Manley, M. E.; Staunton, J. B.; Hermann, R. P.; Frandsen, B. A. Real-Space Visualization of Short-Range Antiferromagnetic Correlations in a Magnetically Enhanced Thermoelectric. Matter 2022, 5, 1–12; https://doi.org/10.1016/j.matt.2022.03.011.Suche in Google Scholar

118. Blume, M. Polarization Effects in the Magnetic Elastic Scattering of Slow Neutrons. Phys. Rev. 1963, 130, 1670; https://doi.org/10.1103/physrev.130.1670.Suche in Google Scholar

119. Maleyev, S. V.; Baryakhtar, V. G.; Suris, P. A. The Scattering of Slow Neutrons by Complex Magnetic Structures. Fiz. Tv. Tela 1962, 4, 3461; Sov. Phys. Solid State 1963, 4, 2533.Suche in Google Scholar

120. Schärpf, O.; Capellmann, H. The XYZ-Difference Method with Polarized Neutrons and the Separation of Coherent, Spin Incoherent, and Magnetic Scattering Cross Sections in a Multidetector. Phys. Status Solidi 1993, 135, 359; https://doi.org/10.1002/pssa.2211350204.Suche in Google Scholar

121. Schweika, W.; Valldor, M.; Reim, J.; Rößler, U. Chiral Spin Liquid Ground State in YBaCo3FeO7. Phys. Rev. X 2022, 12, 021029; https://doi.org/10.1103/physrevx.12.021029.Suche in Google Scholar

122. Halpern, O.; Johnson, M. H. On the Magnetic Scattering of Neutrons. Phys. Rev. 1939, 55, 898; https://doi.org/10.1103/physrev.55.898.Suche in Google Scholar

123. Kibalin, I. A.; Gukasov, A. Local Magnetic Anisotropy by Polarized Neutron Powder Diffraction: Application of Magnetically Induced Preferred Crystallite Orientation. Phys. Rev. Res. 2019, 1, 033100; https://doi.org/10.1103/physrevresearch.1.033100.Suche in Google Scholar

124. Gupta, S. K.; Nielsen, H. H.; Thiel, A. M.; Klahn, E. A.; Feng, E.; Cao, H. B.; Hansen, T. C.; Lelièvre-Berna, E.; Gukasov, A.; Kibalin, I.; Dechert, S.; Demeshko, S.; Overgaard, J.; Meyer, F. Multi-technique Experimental Benchmarking of the Local Magnetic Anisotropy of a Cobalt(II) Single-Ion Magnet. JACS Au 2023, 3, 429–440; https://doi.org/10.1021/jacsau.2c00575.Suche in Google Scholar PubMed PubMed Central

125. Golosovsky, I. V.; Kibalin, I. A.; Gukasov, A.; Roca, A. G.; López-Ortega, A.; Estrader, M.; Vasilakaki, M.; Trohidou, K. N.; Hansen, T. C.; Puente-Orench; Lelièvre-Berna, I. E.; Nogués, J. Elucidating Individual Magnetic Contributions in Bi-magnetic Fe3O4/Mn3O4 Core/Shell Nanoparticles by Polarized Powder Neutron Diffraction. Small Methods 2023, 7, 2201725; https://doi.org/10.1002/smtd.202201725.Suche in Google Scholar PubMed

126. Andersen, K. H.; Argyriou, D. N.; Jackson, A. J.; Houston, J.; Henry, P. F.; Deen, P. P.; Toft-Petersen, R.; Beran, P.; Strobl, M.; Arnold, T.; Wacklin-Knecht, H.; Tsapatsaris, N.; Oksanen, E.; Woracek, R.; Schweika, W.; Mannix, D.; Hiess, A.; Kennedy, S.; Kirstein, O.; Petersson Årsköld, S.; Taylor, J.; Hagen, M. E.; Laszlo, G.; Kanaki, K.; Piscitelli, F.; Khaplanov, A.; Stefanescu, I.; Kittelmann, Th.; Pfeiffer, D.; Hall-Wilton, R.; Lopez, C. I.; Aprigliano, G.; Whitelegg, L.; Moreira, F. Y.; Olsson, M.; Bordallo, H. N.; Martín-Rodríguez, D.; Schneider, H.; Sharp, M.; Hartl, M.; Nagy, G.; Ansell, S.; Pullen, S.; Vickery, A.; Fedrigo, A.; Mezei, F.; Arai, M.; Heenan, R. K.; Halcrow, W.; Turner, D.; Raspino, D.; Orszulik, A.; Cooper, J.; Webb, N.; Galsworthy, P.; Nightingale, J.; Langridge, S.; Elmer, J.; Frielinghaus, H.; Hanslik, R.; Gussen, A.; Jaksch, S.; Engels, R.; Kozielewski, T.; Butterweck, S.; Feygenson, M.; Harbott, P.; Poqué, A.; Schwaab, A.; Lieutenant, K.; Violini, N.; Voigt, J.; Brückel, T.; Koenen, M.; Kämmerling, H.; Babcock, E.; Salhi, Z.; Wischnewski, A.; Heynen, A.; Désert, S.; Jestin, J.; Porcher, F.; Fabrèges, X.; Fabrèges, G.; Annighöfer, B.; Klimko, S.; Dupont, Th.; Robillard, Th.; Goukassov, A.; Longeville, S.; Alba-Simionesco, Ch.; Bourges, Ph.; Guyon Le Bouffy, J.; Lavie, P.; Rodrigues, S.; Calzada, E.; Lerche, M.; Schillinger, B.; Schmakat, Ph.; Schulz, M.; Seifert, M.; Lohstroh, W.; Petry, W.; Neuhaus, J.; Loaiza, L.; Tartaglione, A.; Glavic, A.; Schütz, S.; Stahn, J.; Lehmann, E.; Morgano, M.; Schefer, J.; Filges, U.; Klauser, C.; Niedermayer, C.; Fenske, J.; Nowak, G.; Rouijaa, M.; Siemers, D. J.; Kiehn, R.; Müller, M.; Carlsen, H.; Udby, L.; Lefmann, K.; Birk, J. O.; Holm-Dahlin, S.; Bertelsen, M.; Bengaard Hansen, U.; Olsen, M. A.; Christensen, M.; Iversen, K.; Christensen, N. B.; Rønnow, H. M.; Freeman, P. G.; Hauback, B. C.; Kolevatov, R.; Llamas-Jansa, I.; Orecchini, A.; Sacchetti, F.; Petrillo, C.; Paciaroni, A.; Tozzi, P.; Zanatta, M.; Luna, P.; Herranz, I.; del Moral, O. G.; Huerta, M.; Magán, M.; Mosconi, M.; Abad, E.; Aguilar, J.; Stepanyan, S.; Bakedano, G.; Vivanco, R.; Bustinduy, I.; Sordo, F.; Martínez, J. L.; Lechner, R. E.; Villacorta, F. J.; Šaroun, J.; Lukáš, P.; Markó, M.; Zanetti, M.; Bellissima, S.; del Rosso, L.; Masi, F.; Bovo, C.; Chowdhury, M.; De Bonis, A.; Di Fresco, L.; Scatigno, C.; Parker, S. F.; Fernandez-Alonso, F.; Colognesi, D.; Senesi, R.; Andreani, C.; Gorini, G.; Scionti, G.; Schreyer, A. The Instrument Suite of the European Spallation Source. Nucl. Inst. Methods Phys. Res. A 2020, 957, 163402; https://doi.org/10.1016/j.nima.2020.163402.Suche in Google Scholar

127. Schweika, W.; Violini, N.; Lieutenant, K.; Zendler, C.; Nekrassov, D.; Houben, A.; Jacobs, P.; Henry, P. F. DREAM – A Versatile Powder Diffractometer at the ESS. J. Phys.: Confer. Ser. 2016, 746, 012013; https://doi.org/10.1088/1742-6596/746/1/012013.Suche in Google Scholar

128. Frandsen, B. A.; Baral, R.; Winn, B.; Ovidiu Garlea, V. Magnetic Pair Distribution Function Data Using Polarized Neutrons and ad hoc Corrections. J. Appl. Phys. 2022, 132, 223909; https://doi.org/10.1063/5.0130400.Suche in Google Scholar

129. Shull, C. G.; Strauser, W. A.; Wollan, E. O. Neutron Diffraction by Paramagnetic and Antiferromagnetic Substances. Phys. Rev. 1951, 83, 333; https://doi.org/10.1103/physrev.83.333.Suche in Google Scholar

130. Moseley, D. H.; Taddei, K. M.; Yan, J.; McGuire, M. A.; Calder, S.; Polash, M. M. H.; Vashaee, D.; Zhang, X.; Zhao, H.; Parker, D. S.; Fishman, R. S.; Hermann, R. P. Giant Doping Response of Magnetic Anisotropy in MnTe. Phys. Rev. Mater. 2022, 6, 014404; https://doi.org/10.1103/physrevmaterials.6.014404.Suche in Google Scholar

131. Fennell, T.; Deen, P. P.; Wildes, A. R.; Schmalzl, K.; Prabhakaran, D.; Boothroyd, A. T.; Aldus, R. J.; McMorrow, D. F.; Bramwell, S. T. Magnetic Coulomb Phase in the Spin Ice Ho2Ti2O7. Science 2009, 326, 415; https://doi.org/10.1126/science.1177582.Suche in Google Scholar PubMed

132. Allred, J. M.; Taddei, K. M.; Bugaris, D. E.; Krogstad, M. J.; Lapidus, S. H.; Chung, D. Y.; Claus, H.; Kanatzidis, M. G.; Brown, D. E.; Kang, J.; Fernandes, R. M.; Eremin, I.; Rosenkranz, S.; Chmaissem, O.; Osborn, R. Double-Q Spin-density Wave in Iron Arsenide Superconductors. Nature Phys. 2016, 12, 493–498; https://doi.org/10.1038/nphys3629.Suche in Google Scholar

133. Paddison, J. A. M.; Ehlers, G.; Cairns, A. B.; Gardner, J. S.; Petrenko, O. A.; Butch, N. P.; Khalyavin, D. D.; Manuel, P.; Fischer, H. E.; Zhou, H.; Goodwin, A. L.; Stewart, J. R. Suppressed-moment 2-k Order in the Canonical Frustrated Antiferromagnet Gd2Ti2O7. npj Quantum Mater. 2021, 6, 99; https://doi.org/10.1038/s41535-021-00391-w.Suche in Google Scholar

134. Qureshi, N. Mag2Pol: A Program for the Analysis of Spherical Neutron Polarimetry, Flipping Ratio and Integrated Intensity Data. J. Appl. Crystallogr. 2019, 52, 175–185; https://doi.org/10.1107/s1600576718016084.Suche in Google Scholar

135. Pauling, L. The Structure and Entropy of Ice and of Other Crystals with Some Randomness of Atomic Arrangement. J. Am. Chem. Soc. 1935, 1935, 2680–2684; https://doi.org/10.1021/ja01315a102.Suche in Google Scholar

136. Bernal, J. D.; Fowler, R. H. A Theory of Water and Ionic Solution, with Particular Reference to Hydrogen and Hydroxyl Ions. J. Chem. Phys. 1933, 1, 515; https://doi.org/10.1063/1.1749327.Suche in Google Scholar

137. Hansen, T. C.; Kohlmann, H. Chemical Reactions Followed by in situ Neutron Powder Diffraction. Z. Anorg. Allg. Chem. 2014, 640, 3044–3063; https://doi.org/10.1002/zaac.201400359.Suche in Google Scholar

138. Peterson, V. K.; Auckett, J. E.; Pang, W.-K. Real-Time Powder Diffraction Studies of Energy Materials Under Non-Equilibrium Conditions. IUCrJ 2017, 4, 540–554; https://doi.org/10.1107/s2052252517010363.Suche in Google Scholar

139. Bailey, I. F. A Review of Sample Environments in Neutron Scattering. Z. Kristallogr. 2003, 218, 84–95; https://doi.org/10.1524/zkri.218.2.84.20671.Suche in Google Scholar

140. Klotz, S. Techniques in High Pressure Neutron Scattering; CRC Press: Boca Raton, FL, 2013.10.1201/b13074Suche in Google Scholar

141. Haberl, B.; Guthrie, M.; Boehler, R. Advancing Neutron Diffraction for Accurate Structural Measurement of Light Elements at Megabar Pressures. Sci. Rep. 2023, 13, 4741; https://doi.org/10.1038/s41598-023-31295-3.Suche in Google Scholar PubMed PubMed Central

142. Yang, J.; Muhammad, S.; Jo, M. R.; Kim, H.; Song, K.; Agyeman, D. A.; Kim, Y.-I.; Yoon, W.-S.; Kang, Y.-M. In Situ Analyses for Ion Storage Materials. Chem. Soc. Rev. 2016, 45, 5717–5770; https://doi.org/10.1039/c5cs00734h.Suche in Google Scholar PubMed

143. Finger, R.; Kurtzemann, N.; Hansen, T. C.; Kohlmann, H. Design and Use of a Sapphire Single-Crystal Gas-Pressure Cell for In Situ Neutron Powder Diffraction. J. Appl. Crystallogr. 2021, 54, 839–846; https://doi.org/10.1107/s1600576721002685.Suche in Google Scholar PubMed PubMed Central

144. Bianchini, M.; Suard, E.; Croguennec, L.; Masquelier, C. Li-rich Li1+xMn2–xO4 Spinel Electrode Materials: An Operando Neutron Diffraction Study During Li+ Extraction/Insertion. J. Phys. Chem. C 2014, 118, 25947–25955; https://doi.org/10.1021/jp509027g.Suche in Google Scholar

145. Götze, A.; Stevenson, S. C.; Hansen, T. C.; Kohlmann, H. Hydrogen-Induced Order–Disorder Effects in FePd3. Crystals 2022, 12, 1704; https://doi.org/10.3390/cryst12121704.Suche in Google Scholar

146. Grochala, W.; Hoffmann, R.; Feng, J.; Ashcroft, N. W. The Chemical Imagination at Work in Very Tight Places. Angew. Chem. Int. Ed. 2017, 46, 3620–3642; https://doi.org/10.1002/anie.200602485.Suche in Google Scholar PubMed

147. Besson, J. M.; Nelmes, R. J.; Hamel, G.; Loveday, J. S.; Weill, G.; Hull, S. Neutron Powder Diffraction Above 10 GPa. Physica B 1992, 180–181, 907–910; https://doi.org/10.1016/0921-4526(92)90505-m.Suche in Google Scholar

148. Boehler, R.; Guthrie, M.; Molaison, J. J.; dos Santos, A. M.; Sinogeikin, S.; Machida, S.; Pradhan, N.; Tulk, C. A. Large-Volume Diamond Cells for Neutron Diffraction Above 90 GPa. High Press. Res. 2013, 33, 546–554; https://doi.org/10.1080/08957959.2013.823197.Suche in Google Scholar

149. Haberl, B.; Donnelly, M. E.; Molaison, J. J.; Guthrie, M.; Boehler, R. Methods for Neutron Diffraction Studies on Hydride Superconductors and Other Metal Hydrides. J. Appl. Phys. 2021, 130, 215901; https://doi.org/10.1063/5.0069425.Suche in Google Scholar

150. Sano-Furukawa, A.; Hattori, T.; Arima, H.; Yamada, A.; Tabata, S.; Kondo, M.; Nakamura, A.; Kagi, H.; Yagi, T. Six-Axis Multi-anvil Press for High-Pressure, High-Temperature Neutron Diffraction Experiments. Rev. Sci. Instrum. 2014, 85, 113905; https://doi.org/10.1063/1.4901095.Suche in Google Scholar PubMed

151. Ikuta, D.; Ohtani, E.; Sano-Furukawa, A.; Shibazaki, Y.; Terasaki, T.; Yuan, L.; Hattori, T. Interstitial Hydrogen Atoms in Face-Centered Cubic Iron in the Earth’s Core. Sci. Rep. 2019, 9, 7108; https://doi.org/10.1038/s41598-019-43601-z.Suche in Google Scholar PubMed PubMed Central

152. Benz, S.; Chen, D.; Möller, A.; Hofmann, M.; Schnieders, D.; Dronskowski, R. The Crystal Structure of Carbonic Acid. Inorganics 2022, 10, 132; https://doi.org/10.3390/inorganics10090132.Suche in Google Scholar

153. Grzechnik, A.; Meven, M.; Paulmann, C.; Friese, K. Combined X-ray and Neutron Single-Crystal Diffraction in Diamond Anvil Cells. J. Appl. Crystallogr. 2020, 53, 9–14; https://doi.org/10.1107/s1600576719014201.Suche in Google Scholar PubMed PubMed Central

154. Yamashita, K.; Komatsu, K.; Klotz, S.; Fernández-Díaz, M. T.; Fabelo, O.; Irifune, T.; Sugiyama, K.; Kawamata, T.; Kagi, H. A Nano-polycrystalline Diamond Anvil Cell with Bulk Metallic Glass Cylinder for Single-Crystal Neutron Diffraction. High Press. Res. 2020, 40, 88–95; https://doi.org/10.1080/08957959.2019.1700980.Suche in Google Scholar

155. Terada, N.; Qureshi, N.; Chapon, L. C.; Osakabe, T. Spherical Neutron Polarimetry Under High Pressure for a Multiferroic Delafossite Ferrite. Nat. Commun. 2018, 9, 4368; https://doi.org/10.1038/s41467-018-06737-6.Suche in Google Scholar PubMed PubMed Central

Received: 2024-01-02
Accepted: 2024-03-19
Published Online: 2024-04-29
Published in Print: 2024-06-25

© 2024 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Heruntergeladen am 30.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/zkri-2024-0001/html?lang=de
Button zum nach oben scrollen