research papers
Analysis of diffuse scattering in electron diffraction data for the 32H24Cl2N8O2
determination of Pigment Orange 13, CaUlm University, Central Facility of Electron Microscopy, Materials Science Electron Microscopy, Albert Einstein Allee 11, 89069 Ulm, Germany, bHelmholtz Centre for Infection Research (HZI), SFPR, Inhoffenstraße 7, 38124 Braunschweig, Germany, cHelmholtz Institute for Pharmaceutical Research Saarland (HIPS), MINS, Campus E8.1, 66123 Saarbrücken, Germany, dGoethe University, Institute of Inorganic and Analytical Chemistry, Max-von-Laue-Str. 7, 60438 Frankfurt am Main, Germany, and eSanofi, R&D / PDP / TIDES Analytical Sciences, Building H770, 65926 Frankfurt am Main, Germany
*Correspondence e-mail: tatiana.gorelik@helmholtz-hips.de, m.schmidt@chemie.uni-frankfurt.de
The crystallographic study of two polymorphs of the industrial pyrazolone Pigment Orange 13 (P.O.13) is reported. The β phase was determined using single-crystal X-ray analysis of a tiny needle. The α phase was investigated using three-dimensional electron diffraction. The electron diffraction data contain sharp Bragg reflections and strong diffuse streaks, associated with severe stacking disorder. The structure was solved by careful analysis of the diffuse scattering, and similarities of the unit-cell parameters with the β phase. The structure solution is described in detail and this provides a didactic example of solving molecular crystal structures in the presence of diffuse scattering. Several structural models were constructed and optimized by lattice-energy minimization with dispersion-corrected DFT. A four-layer model was found, which matches the electron diffraction data, including the diffuse scattering, and agrees with X-ray powder data. Additionally, five further phases of P.O.13 are described.
of theKeywords: 3D electron diffraction; electron crystallography; diffuse scattering; stacking disorder; lattice-energy minimization.
CCDC reference: 2160709
1. Introduction
1.1. 3D electron diffraction and stacking disorder
In recent years, electron diffraction (ED) has fast gained popularity, with the development of techniques for three-dimensional (3D) data acquisition and processing (Gruene & Mugnaioli, 2021). 3D ED has been applied to almost all classes of materials, including nanocrystalline organic compounds (Gemmi et al., 2019). Many striking results were obtained for crystal structures, which could not be tackled by other diffraction techniques, such as single crystal X-ray analysis or X-ray powder diffraction. In most cases, the failure of the X-ray methods was associated with (i) small crystal size, (ii) minor amount of material, (iii) polyphasic samples or (iv) severe disorder. The most difficult situations are faced when several of these issues are present simultaneously, e.g. a disordered nanocrystalline non phase-pure sample.
Disorder has a significant influence on the physical properties of materials, and is therefore of fundamental importance for understanding these properties (Tong et al., 2015). Despite being often associated with inorganic materials, disorder is a frequent phenomenon in molecular crystals. In most cases, disorder affects side groups, such as tert-butyl or CF3 (see e.g. Yennawar et al., 2018), or solvent molecules (see e.g. Spek, 2015). Stacking faults are quite frequent in organic crystals, too. Examples include compact molecules such as tris(bicyclo[2.1.1]hexeno)benzene (Bürgi et al., 2005; Schmidt & Glinnemann, 2012), planar aromatic compounds such as the industrial hydrazone pigment Pigment Red 170 (Warshamanage et al., 2014; Teteruk et al., 2014), the pentacyclic pigment αII-quinacridone (Gorelik et al., 2016), and the pyramid-shaped chloro(phthalocyaninato)aluminium (Czech et al., 2017). An unusual example is given by eniluracil, which produces a wide range of disordered structures with significant variability in physical properties, which mimics (Copley et al., 2008). Disorder of supra-molecular columns for compact disk-like molecules have been reported (Schmidt & Neder, 2017; Zehe et al., 2017).
The appeal of stacking disorder lies in the ease of its detection, because any faulted sequence of layers produces streaks of diffuse scattering in ). If such diffuse streaks are observed in a diffraction pattern, there are three possible approaches to treat the data:
(Welberry, 2004(a) Complete neglect of the diffuse scattering in structure solution and leads to the average structure. Generally, the average structure contains a superposition of different possible atomic positions in a For a compound with stacking disorder, the of the average structure is frequently too small to be chemically sensible, as it contains the overlay of two (or more) possible configurations.
(b) Evaluation of the diffuse scattering intensities at the positions of the Bragg reflections leads to an (ordered or disordered) which is generally chemically sensible and provides a good model for the actual structure. Supercells are used to assign side peaks along diffuse streaks as Bragg positions.
(c) Evaluation of the full diffuse scattering provides information on the real structure, including stacking probabilities, preferred local arrangements, deviation from the average structure etc. This approach is tedious, yet has been already applied to several organic compounds (see e.g. Weber & Bürgi, 2002; Welberry, 2004; Bürgi et al., 2005; Weber & Simonov, 2012; Schmidt & Glinnemann, 2012; Teteruk et al., 2014; Welberry & Weber, 2016).
For the evaluation of the diffuse scattering of tiny sub-micron organic crystals, 3D ED is the method of choice. The electron beam can be focused down to very small areas, so even very tiny crystals can be addressed individually. The scattering power of electrons is much higher than that of X-rays, hence, reasonable quality diffraction data can be obtained, despite the small crystal size. There are different flavours of experimental setup for a 3D ED experiment, including selected area ED (SAED), nanodiffraction, static patterns, precession electron diffraction, combined beam tilt/stage tilt, and continuous rotation. An overview of the different methods is given by Gemmi & Lanza (2019) and Gruene & Mugnaioli (2021). Diffraction patterns obtained with precession or continuous rotation represent a physical integration of a certain wedge of the within each frame. The wedge-integrated data are believed to provide a better data quality for structure analysis, compared to static, sequentially collected patterns. A special case is the analysis of electron diffuse scattering (Krysiak et al., 2018, 2020; Mugnaioli & Gorelik, 2019): here, a wedge integration with a relatively large step, either with precession or continuous rotation would smear the data, and the fine details of the intensity distribution within the diffuse streaks would be lost. Therefore, for the analysis of diffuse scattering, static sequentially collected patterns are usually used (Kolb et al., 2019).
In this paper we use 3D ED to determine the α phase of Pigment Orange 13, which shows stacking disorder with strong diffuse scattering.
of the1.2. Pigment Orange 13
The . P.O.13 is an organic hydrazone pigment. Formerly, hydrazone pigments were called `azo pigments', because they were believed to contain an azo moiety (—N=N—C). However, all spectroscopic investigations and single-crystal structure analyses show that these compounds actually adopt the hydrazone-tautomeric form with a —NH–N=C group in the solid state (see e.g. Mustroph, 1987; Whitaker, 1988a,b; Ivashevskaya et al., 2009; Kamei et al., 2011). Hence, they must be named hydrazone pigments instead of azo pigments (Paulus, 1982; Hunger & Schmidt, 2018).
of Pigment Orange 13 (P.O.13) is shown in Fig. 1P.O.13 was invented in 1910 by A. L. Laska in the Chemische Fabrik Griesheim-Elektron in Frankfurt am Main (Laska, 1910, 1911). It has been industrially produced for more than 80 years. Its old name was `Vulcan Orange G extra' (Ershov et al., 1934). Later it was sold, for example, as `Permanent Orange G' by Hoechst and Clariant. P.O.13 is industrially synthesized from 3,3′-dichlorobenzidine and 5-methyl-2-phenyl-3-pyrazolone, see Fig. 2.
Most organic pigments show ). However, no polymorphs were known for P.O.13 until recently. A few years ago, we performed an extensive polymorph screening, and found seven crystal phases (Bekö et al., 2014). The synthesis results in the α phase. The thermodynamically more stable β phase is obtained from the α phase by recrystallization in chlorobenzene, 1,2-dichlorobenzene or 1,2,4-trichlorobenzene. Five other phases (δ, ɛ1, ɛ2, ζ, η) are formed by recrystallization from other solvents, e.g. morpholene, dioxane, 1-chloronaphthalene, or H2SO4 (Bekö et al., 2014).
(Hunger & Schmidt, 2018The α phase is commercially used for printing applications such as printings of packaging. For example, we found P.O.13 in Barilla noodle packaging, and in front covers of Acta Crystallographica Section C and Zeitschrift für Kristallographie, with laser-desorption-ionization P.O.13 is also used as a shading component in printing inks to give yellow pigments a warmer, light orange shade.
Like all pigments, P.O.13 is insoluble in its application medium (e.g. printing ink), being finely dispersed. Particle sizes are typically 50–200 nm. The crystal structures are maintained, and the resulting optical properties strongly depend on the polymorphic form and on the particle size. For example, the β phase of P.O.13 is more opaque and has a slightly more reddish shade than the α phase.
Despite the commercial importance of P.O.13, and despite its long history – P.O.13 is one of the oldest organic pigments – the crystal structures of the compound have never been revealed, hitherto.
In the following, we describe the determination of the β phase of P.O.13 by single-crystal X-ray diffraction. With the knowledge of this structure, we were able to solve the of the commercial α phase of P.O.13 by 3D electron diffraction, including the analysis of the diffuse scattering, supported by lattice-energy minimization. Additionally, we determined the of the ɛ1 phase by single-crystal X-ray diffraction.
of the2. Experimental
2.1. Synthesis and recrystallization
P.O.13 was synthesized by diazotation of 3,3′-dichlorobenzidine and subsequent coupling with 5-methyl-2-phenyl-3-pyrazolone in water, according to Fig. 2, as described by Bekö et al. (2014), resulting in an orange powder of the α phase.
A sample of the α phase with improved crystallinity was obtained by treatment with 2,5-hexanedione. This sample was used for electron diffraction experiments.
The β phase was obtained by the following procedure: P.O.13 (50 mg) was dissolved in boiling chlorobenzene (or 1,2-dichlorobenzene, or 1,2,4-trichlorobenzene) (20 ml), and re-precipitated by slow cooling to room temperature. The precipitate was isolated by filtration and dried at room temperature.
Single crystals of the β phase could be grown following different routes: by recrystallization from e.g. 1,2-dichlorobenzene, 1,2,4-trichlorobenzene or aminoethanol/butanone:
(a) P.O.13 (80 mg) was suspended in aminoethanol (3 ml) using an ultrasound bath. The undissolved portion was removed by filtration. The solution was transferred into a small vial, which was closed using a lid with a pinhole, and put into a flask containing butanone (7 ml) as an anti-solvent. The flask was sealed and kept at room temperature. The butanone slowly diffused via the gas phase into the aminoethanolic solution causing P.O.13 to precipitate as needle-shaped single crystals of the β phase.
(b) P.O.13 (50 mg) was dissolved in boiling 1,2,4-trichlorobenzene (20 ml; b.p. 213°C). After one hour, the solution was allowed to cool slowly to room temperature, resulting in small block-like single crystals of the β phase.
(c) The same as (b), but with 150 mg of P.O.13, giving crystals of the β phase of P.O.13 in the shape of elongated plates.
The γ phase does not exist. The samples which we initially denominated as γ phase turned later out to be mixtures of other phases.
The δ phase was obtained by recrystallization from aminoethanol/dioxane: α-P.O.13 (50 mg) was dissolved in 2-aminoethanol (3 ml). 1,4-Dioxane (7 ml) was slowly added, and the mixture was stored for 5 days at room temperature. The precipitate was isolated by filtration and dried at ambient conditions.
Single crystals of the ɛ1 phase were obtained by crystallization from aminoethanol/dioxane: α-P.O.13 (80 mg) was suspended in aminoethanol (3 ml) using an ultrasound bath for 5 min at room temperature. The remaining solid was removed by filtration. The filtrate was transferred into an open vial, which was placed together with dioxane (7 ml) in a closed flask, and kept at room temperature. The dioxane slowly diffused via the gas phase into the aminoethanol, causing the formation of orange needle-shaped single crystals of the ɛ1 phase with dimensions of about 0.1 mm × 0.01 mm × 0.01 mm.
The ɛ2 phase was obtained by crystallization from morpholene: P.O.13 (60 mg) was heated with morpholene (20 ml; b.p. 129°C) to reflux. The solid was isolated by hot filtration, and dried at room temperature. The resulting powder contained tiny thin needles (probably ɛ2 phase), which were too small for single-crystal X-ray diffraction. The powder pattern showed a mixture of phases β, δ and ɛ2.
The ζ phase was formed by recrystallization of α-P.O.13 (100 mg) in 1-chloronaphthalene (5 ml) at 130°C and slow cooling to room temperature. The pigment was filtrated, and dried at room temperature. A pure ζ phase was obtained only once; reproduction attempts led to mixtures of ζ and β phases, or to the pure β phase.
The amorphous η phase was obtained from concentrated H2SO4: α-P.O.13 (40 mg) was dissolved in concentrated sulfuric acid (3 ml) at room temperature in an open vial. The open vial was placed in a larger vial, and surrounded by water (7 ml). The larger vial was sealed. The water slowly diffused into the H2SO4 solution, causing the pigment to precipitate. The precipitate was filtered, washed with water and dried at room temperature.
2.2. Sample characterization
2.2.1. Methods
All crystalline phases were characterized by powder X-ray diffraction, thermal, spectroscopic and elemental analyses.
X-ray powder patterns were recorded in transmission mode on a Stoe Stadi-P diffractometer equipped with a Ge(111) monochromator and a position-sensitive detector using Cu Kα1 radiation. The sample was rotated during the measurement.
Thermogravimetric analysis (TGA) was carried out using a TGA 92 system (SETERAM Instrumentation). Differential scanning calorimetric (DSC) measurements were performed with a DSC 131 systems (SETERAM Instrumentation) in the temperature window from room temperature up to 500°C.
Spectroscopically the samples were characterized by FTIR using a Shimadzu FTIR-8300 measuring in transmission mode; LDI-TOF-MS using a Voyager-DE STR (Applied Biosystems Inc.) using a nitrogen laser with 337 nm and 10 ns; all liquid 1H NMR studies were carried out using an Avance 250 MHz NMR spectrometer (Bruker) at 300 K in d2-sulfuric acid.
Elemental analyses (EA, only C:H:N quantification) were performed on a vario MICRO cube from Elementar Analytical Systems GmbH.
2.2.2. Analytical data of individual phases
The α phase resulting from synthesis and crystallization experiments shows one single weight loss in the TGA starting at 305°C, and a signal in the DSC resulting from decomposition at approx. 324°C, which points to a solvent-free phase. FTIR and 1H NMR showed the typical signals and the LDI-TOF-MS the typical fragmentation signals at the NH–N bond on one or both sides of the molecule. Furthermore, the FTIR confirmed the bis(hydrazone)-tautomeric form. The solvent-free character could be confirmed by EA calculated for P.O.13 C32H24Cl2N8O2 (%): C 61.64, H 3.88, N 17.97; found: C 61.42, H 3.65, N 18.30.
The β phase exhibits a similar behaviour in thermal analyses as the α phase. The mass loss starts at 305°C in the TGA. The decomposition starts at 311°C, pointing again to a solvent-free phase. In the FTIR and 1H NMR spectra, no significant differences to the α phase could be observed. The EA confirmed the solvent-free form with calculated values for P.O.13 C32H24Cl2N8O2 (%): C 61.64, H 3.88, N 17.97; found: C 59.43, H 3.25, N 16.67.
For the δ phase very similar thermal analytical results could be found as for the α and β phases. Here again a solvent-free phase could be found, which was also confirmed by EA with calculated values of P.O.13 C32H24Cl2N8O2 (%): C 61.64, H 3.88, N 17.97; found: C 60.41, H 3.91, N 17.73.
The ɛ1 and ɛ2 phases occurred only in a mixture with the δ phase. The ɛ1 phase could only be analysed using X-ray powder diffraction and single-crystal X-ray diffraction, the ɛ2 phase only by X-ray powder diffraction.
The ζ phase showed an IR spectrum is similar to that of the other phases. The 1H NMR showed additional peaks, which could not be attributed to P.O.13 or 1-chloronaphthalene. The reason is unknown (by-products from the synthesis?). TGA revealed a mass loss of 4.8% between 120°C and 180°C, which was absent in the other solvent-free phases, and might come from 1-chloronaphthalene (b.p. 260°C) absorbed at the surface of the powder. Melting under decomposition was observed at 321°C.
The η phase showed similar thermal and spectroscopical results as the other solvent-free forms. EA data: calculated for P.O.13 C32H24Cl2N8O2 (%): C 61.64, H 3.88, N 17.97; found: C 59.21, H 3.89, N 17.01. The deviation of calculated and found values may be caused by remaining amounts of water or H2SO4.
2.3. Electron diffraction of the α phase
Samples for TEM investigations were prepared by ultrasonication in n-hexane. A drop of the obtained suspension was placed onto a carbon-coated copper grid (Plano, Wetzlar, S160) and dried in air.
The TEM studies were carried out using a TECNAI F30 TEM (Thermofisher, The Netherlands) equipped with a field emission gun running at 300 kV. Electron diffraction data were collected using an automated acquisition module working in STEM / nano-diffraction modes, as described earlier (Kolb et al., 2007). The beam diameter for the nano-diffraction measurements was 100 nm. Diffraction frames were collected sequentially through the goniometer tilt step of 1°. No electron beam precession was used. Four electron diffraction tilt series were collected with the total tilt range between 50° and 86° (Table 1).
|
The data were processed using EDT Process software (AnaliteX, Sweden), supported by self-written MatLab scripts. Visualization of the reconstructed diffraction volumes were produced using the UCSF Chimera package (Pettersen et al., 2004). Sections of the representing the main crystallographic zones were calculated in PETS2 (Palatinus et al., 2019).
Electron diffraction patterns of the obtained models were kinematically simulated using eMAP (AnaliteX, Sweden) software.
2.4. Single-crystal X-ray structure analyses of the β and ɛ1 phases
2.4.1. β phase
The μS microfocus X-ray source and an APEX2 CCD detector. The structure was solved by with SHELXS97 (Sheldrick, 1990) and refined with SHELXL97 (Sheldrick & Schneider, 1997). All non-H atoms were refined anisotropically.
of a tiny needle was determined using a Bruker SMART three-circle diffractometer equipped with a copper Incoatec IThe structure of the β phase was originally solved with the setting a = 14.425 (4) Å, b = 12.127 (3) Å, c = 33.137 (9) Å, Pbca, Z = 8. To facilitate the comparison of the α and β phases, we transformed the with a′ = c, b′ = a, c′ = b, resulting in the unit-cell parameters a = 33.137 (9) Å, b = 14.425 (4) Å, c = 12.127 (3) Å. The remained to be Pbca, Z = 8. The latter setting is used throughout this paper (Table 2).
|
2.4.2. ɛ1 phase
The ɛ1 phase (dioxane disolvate monohydrate) was determined by single-crystal X-ray diffraction as described for the β phase. The data quality was very limited. The dioxane molecule is severely disordered. The scattering power of the dioxane molecule (C4H8O2, in total 48 electrons) was approximated by eight carbon atoms (48 electrons). The water molecule was approximated by a single oxygen atom.
of the2.5. Generation of ordered structural models
Ordered and disordered structural models were constructed using the software Materials Studio (Version 4.4, BIOVIA Dassault Systèmes, San Diego, USA).
2.6. Lattice-energy minimization
Preliminary lattice-energy minimizations were carried out by force-field methods using the Materials Studio software. The Dreiding force-field (Mayo et al., 1990) was combined with atomic charges calculated by the Gasteiger method (Gasteiger & Marsili, 1980).
The lattice-energy minimizations described in this paper were performed with dispersion-corrected density functional theory (DFT-d) using the program GRACE (Neumann et al., 2008). GRACE uses the VASP code (Kresse & Hafner, 1993; Kresse & Furthmüller, 1996a,b; Kresse & Joubert, 1999) for the DFT calculations. The PBE functional was applied. For the van der Waals interactions, an empirical dispersion correction was used (Neumann & Perrin, 2005).
Two sets of calculations were performed: (a) optimization of the atomic coordinates, with unit-cell parameters fixed to the values determined from electron diffraction, (b) optimization of the unit-cell parameters together with the atomic coordinates.
3. Results and discussion
3.1. of the β phase, serving as the basis for the structure solution of the α phase
Recrystallization of P.O.13 resulted in single crystals with different morphologies: needles, platelets and blocks, see Fig. 3.
All crystals were rather small, so initially all attempts to determine the structure from single-crystal X-ray diffraction failed. Different diffractometers were tried. Finally, the
of a needle with a size of 0.050 mm × 0.050 mm × 0.3 mm could be determined at Sanofi (Frankfurt-Höchst, Germany), using a diffractometer equipped with a microfocus X-ray source and a CCD detector.Despite the different morphologies, all crystals shown in Fig. 3 correspond to the same β phase, as proven by single-crystal X-ray analyses and X-ray powder diffraction.
The β phase crystallizes in the Pbca with unit-cell parameters of a = 33.137 (9), b = 14.425 (4), c = 12.127 (3) Å, α = β = γ = 90°, V = 5797 (3) Å3. Crystallographic data are given in Table 2.
The
analysis proves that P.O.13 possesses the hydrazone tautomeric form, not the azo form, in the solid state – like it was observed for all other industrial organic pigments, too.The β phase is shown in Fig. 4. The two phenyl rings in the centre of the molecule are almost coplanar and form a dihedral angle of φ1 = 9.3°. In the gas phase or in solution, the central biphenyl fragment is twisted by about 40°. Hence, the planarity of the biphenyl fragment is a packing effect, because planar molecules can generally adopt a more efficient packing with a higher packing energy (Schmidt et al., 2007). Several other biphenyl-hydrazone pigments show a planar conformation, too.
of P.O.13 in itsAn isolated molecule of P.O.13 can rotate around the central phenyl–phenyl bond, so it can adopt an overall cis conformation (i.e. with chlorine atoms on the same side of the molecule) or an overall trans conformation (with Cl on opposite sides). Both conformations have a similar intramolecular energy. Astonishingly, the β phase of P.O.13 shows the molecule in an overall cis conformation, see Fig. 4. In contrast, all other known crystal structures of hyrazone pigments based on 3,3′-dichlorobenzidine exhibit the trans conformation, with torsion angles between 151.9 and 180°, see Fig. 5 (Barrow, 2002; Barrow et al., 2000, 2003; Schmidt et al., 2007). The reason why P.O.13 adopts a cis conformation in its β phase is unknown. Apparently, it is a packing effect, but details remains obscure.
The terminal phenyl rings are slightly rotated relative to the adjacent pyrazolone moieties, with torsion angles of 0.5 and −9.3°, respectively. In other pyrazolone pigments, this torsion angle varies between 0° and 35° (see Figs. S1 and S2).
In β-P.O.13, the molecules are stacked on top of each other (Fig. 6). The stacking results in a dense packing with a favorable lattice energy, which would not be possible for twisted molecules.
In the β phase, the contains two layers running parallel to (010). In each layer, the molecules form a herringbone arrangement, see Figs. 6(c) and 6(d). The layers are stacked along the b axis.
3.2. Electron crystallography of the α phase
3.2.1. Crystal morphology
No crystals of the α phase with a size suitable for single-crystal X-ray diffraction could be grown. The X-ray powder diffraction diagram consists of only a few peaks (see Section 3.2.12), which could not be reliably indexed. Therefore, single-crystal 3D ED data were collected.
Electron microscopy revealed that the crystals of α-P.O.13 have an irregular shape with sizes ranging from 200 nm to 1.5 µm. A STEM image of a crystal is shown in Fig. 7. Despite the absence of well defined facets, all studied crystals had the same orientation on the TEM grid. Consequently, one dimension of the crystals (along the electron beam, normal incidence) is apparently significantly shorter than the others. This direction was taken as [100] direction of the crystal structure.
3.2.2. 3D electron diffraction
3D ED datasets of four different crystals were collected. For all datasets, initially an orthorhombic a ∼ 16 Å, b ∼ 14 Å and c ∼ 12 Å was found. The main projections of the reconstructed 3D as well as the main crystallographic zones (obtained by a corresponding cut of the reciprocal space) are shown in Fig. 8.
withThe diffraction patterns of all investigated crystals consist of a mixture of discrete Bragg reflections and strong parallel streaks of diffuse scattering. Such parallel rods of diffuse scattering are typical for crystals with stacking disorder, which are often observed for layered materials (Welberry, 2004).
In layered structures, frequently the interactions between the layers are much weaker than the bonding within a layer. Correspondingly, the crystals grow preferably along the layer directions. In a TEM experiment, in which the crystals are placed onto a flat support normal to the incident electron beam, the layered crystals are usually oriented with their layers parallel to the supporting film, so that the stacking direction is perpendicular to the film, parallel to the incident electron beam. In the presence of stacking faults, or stacking disorder, the associated diffuse scattering usually runs parallel to the electron beam. This situation was, for instance, observed in the electron diffraction investigations of αII-quinacridone (Gorelik et al., 2016). In the case of α-P.O.13, the diffuse scattering is not running parallel to the electron beam, but orthogonal to it. Correspondingly, the stacking direction is not aligned along the shortest crystal dimension, but is oriented perpendicular to it. The explanation for this unusual behaviour is given in Section 3.2.15.
3.2.3. Determination of unit-cell parameters
First, the unit-cell parameters were determined from electron diffraction data, taking into account the positions of the diffuse streaks. The resulting ao, bo and co, in order to distinguish them from other cell settings discussed later.
will be called the original setting of the and denoted with the unit-cell parametersThe unit-cell parameters of the four investigated crystals agreed well, and lead to an orthorhombic . The averaged lattice parameters were ao = 16.1 (3) Å, bo = 14.5 (2) Å, co = 12.3 (1) Å.
within the accepted tolerance for ED data, see Table 13.2.4. Systematic and the space group
The ). The projection of the reciprocal volume along a*, b* and c*, and the zones [100], [010] and [001] are shown in Fig. 8. In the diffraction patterns, the following were observed:
is orthorhombic primitive (Table 1hkl: none
0kl: k = 2n
h0l: l = 2n
hk0: none
h00: not observable, due to the missing cone of reflections along the incident electron beam
0k0: k = 2n
00l: l = 2n.
These Pbc –, which corresponds to the space groups Pbc21 (non-standard setting of Pca21, No. 29), or P 2/b 21/c 21/m (Pbcm, No. 57) (Hahn, 2002). After structure solution (see below), Pbcm could be ruled out, because the molecular arrangement disagrees with the mirror plane.
lead to the3.2.5. Diffuse scattering
The diffuse scattering of α-P.O.13 exhibits three main features:
(i) The diffuse scattering consists of strong diffuse streaks parallel to b* (Fig. 8). Correspondingly, the structure consists of layers parallel to (010), which exhibit a stacking disorder along b. Note that the expression `layer' in the description of a stacking disorder does not need to be a `layer' in the usual meaning for a chemist, but can be any kind of a two-dimensional building block.
(ii) The diffuse streaks are located between the sharp hkl Bragg reflections: layers of reflections with even h consist of sharp Bragg reflections without any diffuse scattering; rows with odd h consist of diffuse scattering only [Figs. 8(a), 8(c), 8(d)]. Correspondingly, the average structure, related to the Bragg reflections only, can be described by a with aav* = 2ao*, hence aav = ao/2 ≃ 8.1 (1) Å, bav = bo, cav = co, αav = βav = γav = 90°. This average structure has translational symmetry with a periodicity of ao/2 ≃ 8.1 (1) Å. If the actual structure is described with the original (ao, bo, co), then each layer can adopt two positions, which differ by Δx = ao/2 ≃ 8.1 (1) Å. The layers themselves are apparently ordered, since there is no indication for any additional diffuse scattering outside the streaks (except thermal diffuse scattering, which is always present).
For the hkl: none; 0kl: k = 2n; h0l: l = 2n), see Fig. 8. Additionally, a new rule appears. In the zone [001] half of the reflections is extinct [marked by diamonds in Fig. 8(d)], corresponding to the reflection condition hk0: h+k = 2n. The combination of these extinction rules leads to the Pbca, corresponding to Pbcn (No. 60), for the average structure. Hence, the average structure has a higher symmetry than the original structure itself, which is a common phenomenon in disordered materials.
of the average structure, all of the original remain valid: ((iii) Within the diffuse streaks, there are intensity maxima at half-integer k values, i.e. at k = ±0.5, ±1.5, ±2.5, ±3.5 etc., see Fig. 8(c). At integer k values, i.e. at the Bragg positions, the intensity approaches minimum. To describe these maxima as Bragg peaks, the reciprocal has to be halved, bL* = bo*/2, see Fig. 9. In real space this enlarges the by 2, corresponding to the lattice parameters of bL = 2bo ≃ 29.1 (4) Å, aL = ao, cL = co, αL = βL = γL = 90°, see Table 3. Hence, the preferred local structure has a repetition unit of bL ≃ 29.1 (4) Å.
‡Unit-cell parameters and atomic coordinates from DFT-d. §Standard setting: P21/c, Z = 4, a = c = 16.6 (3) Å, b = 12.3 (1) Å, β = 58.1 (9)°, α = γ = 90°, V = 2871 Å3. ¶Standard setting: P21/c, Z = 4, a = 15.5174 Å, b = 12.5362, c = 17.0608 Å, β = 59.1173°, α = γ = 90°, V = 2848.27 Å3. |
The large (c) and 9]: hkl: h+k = 2n, which reveals the preferred local structure as being C-centred. Obviously, this extinction rule within the diffuse lines is not strictly followed, hence, the C-centring should be treated as an approximation only.
exhibits the following [Figs. 8In addition to the strong diffuse streaks parallel to b*, the diffraction pattern shown in Fig. 8(f) contains faint diffuse streaks parallel to [201] and . The origin of this feature is discussed in Section 3.2.14.
3.2.6. (two-layer model)
We tried to extract the intensities of electron diffraction reflections and perform an ab initio structure analysis using as implemented in SIR (Burla et al., 2015). The resulting scattering density maps could not be interpreted with a chemically sensible molecular packing of P.O.13, possibly due to the presence of the stacking disorder.
The structure of α-P.O.13 was solved by a manual approach. Such a procedure was common in the early days of X-ray crystallography. For example, Kathleen Lonsdale solved the triclinic structure of hexamethylbenzene in 1929 by careful visual consideration of the distribution of individual reflection intensities in the diffraction pattern (Lonsdale, 1929). Also, in the early days of electron crystallography on organic compounds, a very similar approach was used, e.g., for the structure solution of Pigment Red 53:2 (Gorelik et al., 2009). Still today, the manual approach is valuable, e.g., for difficult structure solutions from powder data, or for quasicrystals, see Fig. 10.
The number of molecules per α and β phases, revealing that the of the α phase (with lattice parameters ao, bo, co) contains 4 molecules of P.O.13.
was deduced from the volumes of theThe α and β phases of P.O.13 have very similar lattice parameters, with aoα ≃ ½ aβ, boα ≃ bβ, coα ≃ cβ, α = β = γ = 90°. Also the space groups showed some similarities (Pbca for the α phase, Pbc21 or Pbcm for the β phase in its original unit-cell setting). These observations indicated that the of the α phase might contain somehow similar features as the structure of the β phase.
For the structure solution of α-P.O.13, we assumed that neighbouring molecules are arranged in a similar way as in the β phase (Fig. 6). The structure was then solved manually by three of us (JT, MUS and TEG) using print-outs of the molecular packing shown in Figs. 6(b)–6(d).
To account for the reduction of the unit-cell parameter a0 from 33.1 to 16.1 Å, we omitted half of the content of the β phase, see Fig. 11.
The obtained structure consists of two layers in the b direction, like the β phase. The lattice parameter a0α is now a0α = ½ aβ, as seen from the ED data. Upon this cell reduction, all symmetry elements of the β phase, which connected the omitted half of the with the remaining one, were lost. From the original symmetry P 21/b 21/c 21/a (Fig. 4) only Pbc21 remains. No additional symmetry elements were generated, apart from the translation vector of ½aβ (which also leads to a doubled density of 21 and b symmetry elements). Consequently, the new structure has Pbc21 symmetry, see Fig. 12(a). This symmetry fully matches the extinction conditions observed in the ED patterns. Pbc21 is a non-standard setting of Pca21 (space group No. 29).
Upon the cell reduction from the β phase to the α phase, the symmetry of a single molecular layer changes from P21ca [Fig. 6(c)] to P12/c1, see Fig. 12(b). However, the twofold axis and the inversion centre are only elements. Only the c-glide plane is a global one.
This structural model is ordered (ordered two-layer model). It shows a sensible molecular geometry, reliable intermolecular contacts, and a dense molecular packing. The model was validated by lattice-energy minimization (see below). Hence, from the crystal-chemical point of view, this model is fully sensible. The unit-cell parameters from DFT-d optimization were similar to those derived from ED. The extinction conditions are fulfilled as well. However, the model does not include the stacking disorder and yet, and does not give any explanation for the diffuse scattering.
3.2.7. Stacking disorder
In the ordered two-layer model, the molecules in the first layer (layer A) are centred at x0 = , in the second layer (layer B) at x0 = , see Figs. 12(b) and 12(c). Hence, the centre of the second layer is shifted against the centre of the first layer by a/4. The diffuse scattering revealed that every layer has two possible positions, which differ by ± a/2. Consequently, the first layer could alternatively be located at x0 = (layer C), and the second layer at x0 = (layer D), see Fig. 13. In other words, each subsequent layer is shifted in respect to the preceding either by ± a/4. This ambiguity leads to the stacking disorder.
The superposition of both lateral positions in each layer leads to the average structure. In accordance with the ED data, this average structure can be described with a smaller aav = ao/2, bav = bo , cav = co. The average structure has the Pbcn, see Fig. 13. The and the systematic are, again, in perfect agreement with the electron diffraction data.
withWithin each layer, all molecules must have the same lateral shift, i.e. all molecules in the first layer must either all be at position A, or all at position C, and in the second layer either all at B or all at D. A mixture of molecular positions within a layer would result in large voids or severe molecular overlap. Hence, the layer itself is ordered, which explains the absence of diffuse scattering in other directions apart from b*.
There is an infinite number of possible stacking sequences, periodic and non-periodic. The simplest periodic stacking sequences are |AB|AB|AB| (ordered two-layer model), |ABCB|ABCB|, and |ABCD|ABCD| (the vertical lines denote the repeating unit). A stacking sequence with a periodicity unit of ten layers (|ABADABADAB|ABADABADAB|) is shown in Fig. 14.
3.2.8. Preferred local structure (four-layer model)
The analysis of the intensity distribution within the ED data revealed that the preferred local structure has a periodicity in stacking direction of bL = 2b0 (§3.2.5). Hence, the corresponding model should consist of four layers. Only two symmetrically different four-layer sequences are possible: |ABCD|ABCD| and |ABCB|ABCB|. The ED data clearly showed a C-centred pattern; consequently, the sequence must be |ABCD|. This structure has a C1121/g, in which g stands for a glide plane with a translational vector of (a + b)/4 (Fischer & Koch, 2011)1. C1121/g is an unconventional setting of P21/c [aP21/c = (−aL + bL)/2, bP21/c = cL, cP21/c = (aL + bL)/2]. This unconventional setting was chosen in order to describe the model with the same unit-cell parameters as the two-layer model, except for a doubling of the b axis. The structure of the four-layer model is shown in Fig. 15(a).
The systematic 1 axis agree with the ED data. The for the g-glide plane are hk0: h + k = 4n for |ABCD|, and hk0: h − k = 4n for its mirror image |DCBA|. These are not well visible in the diffraction pattern, because they are buried in the of the C centring of the requiring hk0: h + k = 2n.
for the 2The two-layer model as well as the four-layer model consist of molecules in the cis conformation. We tried to build similar models with molecules in trans conformation. However, the steric requirement of trans molecules considerably differs from that of cis-molecules. In all models built with the trans conformation, the molecules showed unreliably short intermolecular contacts. Lattice-energy minimization lead to an enlargement of the resulting in unit-cell parameters, which strongly deviated from the values obtained by ED. Hence, the molecular packing in α-P.O.13 is possible only for molecules in cis conformation.
3.2.9. Structure refinement
We made a fast attempt to refine the different structural models against ED data using the least-squares kinematical
yet soon abandoned this idea. The quality of the data was not sufficient for a quantitative treatment of the reflection intensities, possibly due to the static data collection procedure and the presence of diffuse scattering, which is known to deteriorate the data quality.Due to the small crystal size and the presence of diffuse scattering, the
against X-ray powder data was not reliable, either (see below).Therefore, the atomic coordinates were optimized by quantum-mechanical methods using density-functional theory with dispersion correction (DFT-d). These calculations were performed on the ordered two-layer model and the ordered four-layer model. The average structure could not be treated by quantum-mechanical methods, due to the disorder with overlapping atoms and an occupancy of 0.5 for all atoms. In the DFT-d calculations of the ordered models, the unit-cell parameters were fixed to the values obtained from ED. Upon optimization, the structures changed only slightly, which proves that the structural models are crystallochemically sensible.
3.2.10. Comparison of simulated and experimental electron diffraction patterns
For the structural models with two, four and ten layers, electron diffraction pattern of the [001] zone, which comprises the diffuse scattering rows, were simulated (Fig. 16). The simulated patterns were compared to the [001] section, extracted from the experimental volume [Fig. 8(d)]. Fig. 16 shows the comparison of simulated and experimental patterns. The strong Bragg reflections of the average structure are correctly reproduced by all three models (some of the Bragg reflections are marked by dashed circles in Fig. 16.) Note that the [001] zone of the four-layer model |ABCD| has only twofold rotation symmetry, whereas the two-and ten-layer models have 2mm symmetry, like the experimental pattern. For a better comparison, the mm symmetry was added to the four-layer model, corresponding to a mixture of |ABCD| and |DCBA| sequences, which is a quite reasonable assumption for a real crystal.
With increasing numbers of layers in the model, the diffuse rows are developing. The experimental distribution of intensities along the diffuse lines follows a certain rule: for the rods at ± 1k0 rod, diffuse intensity is mainly concentrated around the positions with k = 1.5, 3.5, 5.5, and 6.5 [see small diamonds in Figs. 16(a)–16(c)]. This distribution cannot be achieved with a two-layer model [Fig. 16(a)], which wrongly simulates the intensities at integer k values. Correspondingly, a regular AB stacking is not a proper representation of the structure. In contrast, the experimental intensity maxima in the ± 1k0 rod match the simulated intensities of the four-layer model. In addition, the intensities on the diffuse streaks at ± 3k0 and ± 5k0 are reproduced quite well. [Fig. 6(b)]. Hence, the four-layer model describes the diffuse scattering much better. Apparently, the |ABCD| (or |DCBA|) stacking motif is the most prominent one in the structure. The ten-layer model demonstrates that a larger periodicity of the stacking sequence leads to the formation of extended diffuse streaks, like in the experimental pattern.
3.2.12. Structure validation by X-ray powder diffraction
The different structural models were validated by lattice-energy minimizations with DFT-d, and by comparison with the experimental powder pattern.
Two sets of DFT-d optimizations were performed, one with unit-cell parameters fixed to the values obtained by ED, the other one with free unit-cell parameters.
In the DFT-d calculations with unit-cell parameters from ED, the four-layer model is by 0.74 kJ mol−1 more favourable than the two-layer model. In the DFT-d optimizations with free unit-cell parameters, the energy difference increases to 1.86 kJ mol−1. Both values point to a statistical disorder with a slight preference for a local stacking with a sequence |ABCD| (or |DCBA|, respectively) over |AB|AB|.
Upon optimization with free unit-cell parameters, the two-layer structure changed only slightly, and remained orthorhombic. In the four-layer model, the symmetry (C1121/g) was maintained, but the angle γ changed from 90° to 83.69°, see Table 3 and Fig. 15(b). This change corresponds to a lateral shift of the layers by 0.78 Å per layer. Such a lateral shift is well possible in a layer structure. In a real crystal, the lateral shift `to the right' within an |ABCD| domain would be compensated by a corresponding shift `to the left' in a |DCBA| domain. Both, |ABCD| and |DCBA| domains are equally frequent, so that the overall remains orthorhombic. The small energy difference between |AB|AB|, |ABCD| and |DBCA| indicates that the domains with a strict |ABCD| or |DCBA| stacking are actually quite small.
Large domains with a strict |ABCD| or |DCBA| stacking would be visible in diffraction patterns, because the local change of the γ angle from 90° to 83.69° (for |ABCD|) or 96.31° (for |DCBA|) would be visible as a splitting of the Bragg peaks. Such a splitting was not observed, neither in the ED patterns, nor in the X-ray powder patterns. Hence, the diffraction data confirm that the ordered domains with a strict |ABCD| (or |DCBA|) sequence cannot be very large.
In principle, P.O.13 could form a structure, which consists only of the lowest-energy stacking sequence |ABCD|. This structure would be a different polymorph. Since the unit-cell parameters, especially the angle γ, differ from that of the α phase, its powder pattern would significantly differ from the pattern of the α phase. Hence, this polymorph would be easily recognisable from its X-ray powder diffraction pattern. However, in the hundreds of powder patterns, which we recorded during the polymorph screening and in recrystallization attempts, we never observed the formation of this phase.
3.2.11. Structure validation by X-ray powder diffraction
X-ray powder patterns were calculated for the two-layer model, the four-layer model, the ten-layer model, and the average structure. All powder patterns are very similar, and match quite well the experimental powder pattern of the α phase, see Fig. 17. The main differences were found in the region below 12° 2θ. Here, the simulated powder patterns show reflections, depending on the of the structural model. The experimental powder pattern does not show any of these reflections, pointing to a statistical disorder without large ordered domains. Hence, the overall is confirmed, but details on the preferred stacking sequence cannot be derived.
3.2.13. of the α phase
The final structure was obtained from the combination of experimental electron diffraction data and the DFT-d calculations. Unit-cell parameters and the overall molecular packing were obtained from ED, whereas precise atomic coordinates originate from the DFT-d calculations. These structural models, the ordered two-layer and the ordered four-layer model, should be regarded as the `final' structural models for the α phase of P.O.13. The two models have been deposited at the Cambridge Structural Database under the reference codes 2160710 (two-layer) and 2160709 (four-layer). The structures are shown in Figs. 12 and 15(a).
The α phase of P.O.13 is disordered. The ordered two-layer and four-layer models give a good representation of the structure, concerning the unit-cell parameters, the molecular geometry and the arrangement of neighbouring molecules.
of theThe two-layer model is the simplest ordered model. It is crystallochemically sensible, but does not reproduce the positions of the intensity maxima in diffuse scattering. These maxima are reproduced much better by the four-layer model. Hence, the four-layer model gives a much better representation of the actual local structure.
In both models, the molecular geometry is very similar to that in the β phase, see Fig. 18. The molecules adopt an overall cis conformation, like in the β phase.
The π–π stacking of molecules along the c-axis is similar to that in the β phase, but neighbouring stacks within a layer adopt a parallel packing instead of a herringbone packing, see Fig. 12.
3.2.14. Why is the α phase disordered?
Simply spoken, the α phase is disordered, because the different stacking possibilities have similar lattice energies, see §3.2.11.
For a crystallochemical explanation of the disorder, let us consider one ordered layer, see Fig. 19(a). The is that of the ordered two-layer model. If this layer is shifted by Δx = 0.5, then the molecule A overlaps with the molecules C1 and C2 (in yellow). The overlapping three molecules are shown in Figs. 19(b) and 19(c). The mutual arrangement of these three molecules can also be described through a shift of the molecule by half a molecular length along the long axis of the molecule, see Figs. 19(b) and 19(c).
The molecule has the shape of a flat caterpillar. On the lower side, the methyl groups stick out. On the upper side, the chlorine atoms and the phenyl groups stick out as bumps. A shift of the molecule along its long molecular axis by half a molecular length leads to exactly the same positions of the methyl groups at the lower side, whereas at the upper side the chlorine bumps are replaced by the phenyl bumps. Hence, the combination of molecules C1 and C2 reproduces the shape of molecule A almost exactly. In other words: the chain of molecules along their long direction has a shape with a periodicity of half a molecular length. This `pseudosymmetrical' molecular shape (Hörnig et al., 1993) allows the layers to be shifted by Δx = 0.5, with almost no change in the shape of the surface of the layer. The layer surface has a periodicity of a/2 instead of a. Since there are no hydrogen bonds and no strong electrostatic interactions between the layers, any subsequent layer can adopt two lateral positions which differ by a/2.
Figs. 19(a) and 19(b) suggest that an entire chain of molecules [molecules C1 + C2 in Fig. 19(a)] might be shifted along the chain direction by half a molecule, without disturbing the interactions with the neighbouring layers. However, such a chain shift would change the interactions to neighbouring chains within the layer [e.g., the van der Waals and Coulomb interaction of molecule C2 with its neighbours E and F in Fig. 19(a)]. This shift corresponds to a stacking disorder of the chains within the layer. Such a disorder would cause diffuse streaks parallel to [201] and . Indeed, the experimental [010] zone pattern [Fig. 8(f)] contains faint diffuse streaks in these directions.
Why does the β phase not show any disorder? The reason is apparent from Fig. 19(a): in the β phase, molecule C1 has a different orientation [see Fig. 6(c)], hence the layer does not have a pseudosymmetric surface like in the α phase.
3.2.15. On the morphology of the α phase
In the α phase the weakest intermolecular interactions are in the a direction. There are only van der Waals interactions between the terminal phenyl rings, see Fig. 19(a). Correspondingly, the crystal morphology is a platelet parallel to (100), which explains the observation from TEM and ED.
of the3.3. Further crystal phases
In total, seven crystal phases are known for P.O.13 (Bekö et al., 2014). The α phase is formed directly in the synthesis. The β phase, which is the thermodynamically more stable, appears after recrystallization from most solvents. The β phase is also formed if P.O.13 is dissolved in aminoethanol and precipitated with methanol, ethyl acetate, acetone, butanone or toluene. Crystallization from 2-aminoethanol/1,4-dioxane at room temperature leads to two different phases, δ or ɛ1, depending on the crystallization conditions. When a solution of P.O.13 in 2-aminoethanol is treated with liquid dioxane, the δ phase is obtained, which is a solvate. In contrast, adding the dioxane slowly via gas phase diffusion leads to the ɛ1 phase, which turned out to be a dioxane solvate hydrate. The ɛ2 phase is formed upon recrystallization in morpholene. It appears to be a morpholene solvate. It could be obtained only as mixtures with the δ and/or β phases. The ζ phase is obtained by recrystallization in 1-chloronaphthalene. The investigated powder of the ζ phase probably contains traces of other phases. An amorphous phase of P.O.13, called η-phase is formed by treating P.O.13 with concentrated sulfuric acid, followed by precipitation with water vapours.
The phases α, β, δ and η are solvent-free, according to TGA, DSC, IR and elemental analyses. Phases ɛ1 and ɛ2 are solvates. The composition of the ζ phase could not be determined (Probably the sample was not pure).
The powder X-ray diffraction patterns of all phases are shown in Fig. 20.
The low crystallinity of the phases δ and ζ prevented the determination of their crystal structures. The η phase is amorphous.
6.3. of the phases ɛ1 and ɛ2
The phase ɛ1 is a solvate of P.O.13 with dioxane and water in a ratio of P.O.13: dioxane: H2O = 1:2:1. The of this phase was determined by single-crystal X-ray analysis (see Table 2). The crystal was a thin needle. The final R values were poor, mainly due to heavily disordered solvent molecules. The compound crystallizes in the monoclinic C2/c, Z = 4, Z′ = ½. The P.O.13 molecule and the water molecule are situated on crystallographic twofold axes, the dioxane molecule in the general position. Surprisingly, this phase of P.O.13 also adopts a cis conformation, like in the α and β phases, in contrast to all other biphenyl-hydrazone pigments. The dihedral angle φ1 of the central biphenyl moiety is 22.4°. Despite this deviation from planarity, the π–π stacking of neighbouring molecules is very similar to those in α and β phases.
The terminal phenyl rings are almost coplanar with the pyrazole rings (φ2 = φ3 = 1.22°).
The molecules arrange in layers. The voids in the layers are filled by dioxane and water molecules. The dioxane molecules are orientationally disordered. The water molecule donates hydrogen bonds to two dioxane molecules. The P.O.13 molecule forms only intramolecular hydrogen bonds, but no intermolecular hydrogen bond to the dioxane or water molecules. The packing is shown in Fig. 21.
The ɛ2 phase is a morpholene disolvate. It could not be obtained as a pure phase, but only as a mixture with the phases β and/or δ. Single crystals could not be grown. However, the high similarity of the powder patterns of the ɛ1 and ɛ2 phases suggests that the of the ɛ2 phase is very similar to that of the ɛ1 phase. Apparently, the two dioxane molecules are just replaced by two morpholene molecules, which is easily possible, because the morpholene molecule deviates from a dioxane molecule only by an exchange of O versus NH; hence their steric requirements are similar. Whether the morpholene disolvate again contains a water molecule, remains unknown.
4. Conclusion
The industrially relevant α phase of P.O.13, which is used to print covers of Acta Crystallographica Section C, is a nanocrystalline powder with severe stacking disorder. All attempts to improve the crystallinity and to solve the structure by powder or single-crystal X-ray diffraction had failed. We, therefore, turned to electron diffraction. The electron diffraction data contained intense diffuse scattering, which prevented a classical structure solution. Yet, a careful analysis of the positions of the Bragg reflection and of the diffuse scattering, and a certain similarity of the unit-cell parameters of the α phase with the previously determined β phase allowed us to solve the with paper and pencil. Different structural models were built, consisting of two, four and ten layers. These models were subsequently validated by lattice-energy minimization with DFT-d. The structural model with four layers gave a quite good fit to the experimental electron diffraction data, including the main features of the diffuse scattering.
P.O.13 is the first diaryl pigment which exhibits a cis conformation of the central biphenyl fragment. This conformation is found in all four phases (α, β, ɛ1 and ɛ2), although the individual molecule can adopt cis or trans conformations with similar energies, and all other diaryl pigments adopt the trans conformation in the solid state.
We finally would like to emphasize the power of non-standard approaches for
analysis, which are rarely used nowadays with the availability of automated and standardized procedures. Yet manual approaches sometimes can be the only choice for complex crystallographic problems.The most difficult problems are the most interesting ones.
Supporting information
CCDC reference: 2160709
Pigment_Orange_13_alpha phase_stacking_disorder_P21c-model. DOI: https://doi.org/10.1107/S2052520623000720/je5050sup1.cif
Pigment_Orange_13_alpha phase_stacking_disorder_P21c-model. DOI: https://doi.org/10.1107/S2052520623000720/je5050Isup2.hkl
Supporting information file. DOI: https://doi.org/10.1107/S2052520623000720/je5050sup3.pdf
Supporting information file. DOI: https://doi.org/10.1107/S2052520623000720/je5050Isup4.cml
Monoclinic, P2_1/c | Z = 4 |
a = 16.6 (3) Å | T = 293 K |
b = 12.3 (1) Å | Irregular, orange |
c = 16.6 (3) Å | 0.0002 × 0.0002 × 0.0001 mm |
β = 58.1 (9)° |
Experimental. alpha-phase. Electron diffraction. Stacking disorder. 4-layer model in P21/c. The structure was solved by electron diffraction, and subsequently opmimised by lattice-energy minimisation with fixed lattice parameters, using dispersion-corrected density-functional theory with the program GRACE (PBE potential with dispersion correction according to Neumann & Perrin (2005)). The diffraction pattern shows strong diffus scattering as streaks in 101 direction. The structure has a stacking disorder along [101]. The data given here represent an ordered 4-layer model, which provides a reliable molecular geometry and a reliable arrangement of neighbouring molecules, and explains the position of the maxima of the diffuse scattering For a more simple model of the stacking, see the 2-layer model of the alpha phase in Pca21. |
Geometry. The structure was solved by electron diffraction, and subsequently opmimised by lattice-energy minimisation with fixed lattice parameters from electron diffraction, using density-functional theory with the program GRACE. PBE-Potential with dispersion correction according to Neumann & Perrin (2005). Hence, the molecular geometry reflects the DFT-d geometry. |
x | y | z | Biso*/Beq | ||
Cl1 | −1.016323 | 0.981293 | 1.241800 | ||
C34 | −1.038231 | 1.046552 | 1.162792 | ||
C33 | −1.107621 | 1.004060 | 1.147204 | ||
C35 | −0.986717 | 1.139088 | 1.115493 | ||
N8 | −1.157516 | 0.910742 | 1.193706 | ||
C40 | −1.125023 | 1.058868 | 1.083806 | ||
H36 | −0.932229 | 1.166585 | 1.128680 | ||
C37 | −1.004309 | 1.194561 | 1.052108 | ||
N7 | −1.215748 | 0.866122 | 1.172090 | ||
H9 | −1.150436 | 0.874494 | 1.246157 | ||
C38 | −1.075091 | 1.151929 | 1.038269 | ||
H41 | −1.179389 | 1.027185 | 1.072361 | ||
C42 | −0.952386 | 1.293917 | 1.002098 | ||
C27 | −1.260580 | 0.777913 | 1.219304 | ||
H39 | −1.093458 | 1.193218 | 0.992163 | ||
C43 | −0.960036 | 1.336678 | 0.927500 | ||
C49 | −0.892905 | 1.349068 | 1.025174 | ||
C26 | −1.253088 | 0.722670 | 1.293362 | ||
C28 | −1.325910 | 0.715076 | 1.206133 | ||
H44 | −1.002260 | 1.295322 | 0.904282 | ||
C45 | −0.912563 | 1.429880 | 0.879512 | ||
C48 | −0.844214 | 1.442486 | 0.976376 | ||
H50 | −0.883072 | 1.320361 | 1.081552 | ||
O3 | −1.202343 | 0.749887 | 1.325472 | ||
N5 | −1.313588 | 0.634567 | 1.317379 | ||
N6 | −1.356439 | 0.630923 | 1.263432 | ||
C29 | −1.356649 | 0.738553 | 1.138612 | ||
H46 | −0.918183 | 1.460883 | 0.821325 | ||
C47 | −0.854470 | 1.485468 | 0.903657 | ||
Cl2 | −0.770502 | 1.506658 | 1.005676 | ||
C25 | −1.336501 | 0.553064 | 1.385969 | ||
H30 | −1.400460 | 0.672347 | 1.138164 | ||
H31 | −1.294526 | 0.749225 | 1.066472 | ||
H32 | −1.397498 | 0.814706 | 1.158193 | ||
N10 | −0.807660 | 1.579980 | 0.856134 | ||
C15 | −1.391637 | 0.463859 | 1.390827 | ||
C23 | −1.306026 | 0.561553 | 1.450224 | ||
H11 | −0.758732 | 1.617534 | 0.867775 | ||
N12 | −0.824723 | 1.622954 | 0.794049 | ||
H16 | −1.415781 | 0.458696 | 1.341485 | ||
C17 | −1.416116 | 0.384418 | 1.459748 | ||
C21 | −1.332285 | 0.481637 | 1.519416 | ||
H24 | −1.263143 | 0.630666 | 1.445855 | ||
C51 | −0.777714 | 1.711447 | 0.749891 | ||
H18 | −1.459258 | 0.315783 | 1.462883 | ||
C19 | −1.387163 | 0.392762 | 1.524727 | ||
H22 | −1.310711 | 0.489986 | 1.570862 | ||
C52 | −0.785691 | 1.769900 | 0.679353 | ||
C57 | −0.707280 | 1.769914 | 0.761148 | ||
H20 | −1.407490 | 0.331440 | 1.579481 | ||
N13 | −0.729050 | 1.854534 | 0.649450 | ||
C53 | −0.846660 | 1.741220 | 0.641764 | ||
O4 | −0.679308 | 1.745946 | 0.815866 | ||
N14 | −0.680950 | 1.856192 | 0.698219 | ||
H54 | −0.833266 | 1.797333 | 0.584729 | ||
H55 | −0.922200 | 1.744977 | 0.698489 | ||
H56 | −0.831925 | 1.657499 | 0.614523 | ||
C58 | −0.615611 | 1.940796 | 0.678533 | ||
C59 | −0.605393 | 2.024772 | 0.616500 | ||
C67 | −0.560399 | 1.940918 | 0.719858 | ||
H60 | −0.648241 | 2.023969 | 0.584657 | ||
C61 | −0.539545 | 2.107149 | 0.595602 | ||
C65 | −0.494628 | 2.024000 | 0.697612 | ||
H68 | −0.568358 | 1.875660 | 0.767663 | ||
H62 | −0.531660 | 2.171277 | 0.546570 | ||
C63 | −0.483352 | 2.107221 | 0.635308 | ||
H66 | −0.450525 | 2.021375 | 0.728291 | ||
H64 | −0.430610 | 2.170372 | 0.617890 |
Cl1—C34 | 1.7311 | C25—C15 | 1.4031 |
C34—C33 | 1.4068 | C25—C23 | 1.4041 |
C34—C35 | 1.3900 | N10—H11 | 1.0367 |
C33—N8 | 1.3864 | N10—N12 | 1.3145 |
C33—C40 | 1.4015 | C15—H16 | 1.0865 |
C35—H36 | 1.0899 | C15—C17 | 1.3912 |
C35—C37 | 1.4072 | C23—C21 | 1.3943 |
N8—N7 | 1.3148 | C23—H24 | 1.0859 |
N8—H9 | 1.0379 | N12—C51 | 1.3134 |
C40—C38 | 1.3802 | C17—H18 | 1.0899 |
C40—H41 | 1.0871 | C17—C19 | 1.3963 |
C37—C38 | 1.4109 | C21—C19 | 1.3951 |
C37—C42 | 1.4723 | C21—H22 | 1.0906 |
N7—C27 | 1.3143 | C51—C52 | 1.4376 |
C38—H39 | 1.0871 | C51—C57 | 1.4653 |
C42—C43 | 1.4104 | C19—H20 | 1.0878 |
C42—C49 | 1.4068 | C52—N13 | 1.3112 |
C27—C26 | 1.4653 | C52—C53 | 1.4840 |
C27—C28 | 1.4385 | C57—O4 | 1.2533 |
C43—H44 | 1.0875 | C57—N14 | 1.3871 |
C43—C45 | 1.3801 | N13—N14 | 1.4048 |
C49—C48 | 1.3915 | C53—H54 | 1.0946 |
C49—H50 | 1.0894 | C53—H55 | 1.1001 |
C26—O3 | 1.2561 | C53—H56 | 1.0988 |
C26—N5 | 1.3846 | N14—C58 | 1.4105 |
C28—N6 | 1.3122 | C58—C59 | 1.4034 |
C28—C29 | 1.4836 | C58—C67 | 1.4042 |
C45—H46 | 1.0865 | C59—H60 | 1.0867 |
C45—C47 | 1.4015 | C59—C61 | 1.3925 |
C48—C47 | 1.4071 | C67—C65 | 1.3945 |
C48—Cl2 | 1.7268 | C67—H68 | 1.0861 |
N5—N6 | 1.4071 | C61—H62 | 1.0900 |
N5—C25 | 1.4093 | C61—C63 | 1.3969 |
C29—H30 | 1.0940 | C65—C63 | 1.3950 |
C29—H31 | 1.0984 | C65—H66 | 1.0902 |
C29—H32 | 1.0996 | C63—H64 | 1.0884 |
C47—N10 | 1.3886 | ||
Cl1—C34—C33 | 119.10 | N5—C25—C15 | 119.47 |
Cl1—C34—C35 | 119.64 | N5—C25—C23 | 120.53 |
C33—C34—C35 | 121.26 | C15—C25—C23 | 119.99 |
C34—C33—N8 | 120.73 | C47—N10—H11 | 121.67 |
C34—C33—C40 | 118.18 | C47—N10—N12 | 119.60 |
N8—C33—C40 | 121.09 | H11—N10—N12 | 118.73 |
C34—C35—H36 | 117.92 | C25—C15—H16 | 119.41 |
C34—C35—C37 | 120.68 | C25—C15—C17 | 119.65 |
H36—C35—C37 | 121.39 | H16—C15—C17 | 120.94 |
C33—N8—N7 | 120.16 | C25—C23—C21 | 119.41 |
C33—N8—H9 | 121.54 | C25—C23—H24 | 119.57 |
N7—N8—H9 | 118.31 | C21—C23—H24 | 121.02 |
C33—C40—C38 | 120.49 | N10—N12—C51 | 117.38 |
C33—C40—H41 | 118.52 | C15—C17—H18 | 118.83 |
C38—C40—H41 | 120.98 | C15—C17—C19 | 120.83 |
C35—C37—C38 | 117.41 | H18—C17—C19 | 120.34 |
C35—C37—C42 | 122.19 | C23—C21—C19 | 120.92 |
C38—C37—C42 | 120.40 | C23—C21—H22 | 119.25 |
N8—N7—C27 | 116.60 | C19—C21—H22 | 119.82 |
C40—C38—C37 | 121.96 | N12—C51—C52 | 125.82 |
C40—C38—H39 | 117.52 | N12—C51—C57 | 127.87 |
C37—C38—H39 | 120.50 | C52—C51—C57 | 106.31 |
C37—C42—C43 | 120.44 | C17—C19—C21 | 119.21 |
C37—C42—C49 | 122.17 | C17—C19—H20 | 120.80 |
C43—C42—C49 | 117.39 | C21—C19—H20 | 119.99 |
N7—C27—C26 | 127.60 | C51—C52—N13 | 110.43 |
N7—C27—C28 | 125.96 | C51—C52—C53 | 126.82 |
C26—C27—C28 | 106.42 | N13—C52—C53 | 122.74 |
C42—C43—H44 | 120.64 | C51—C57—O4 | 127.02 |
C42—C43—C45 | 121.81 | C51—C57—N14 | 103.57 |
H44—C43—C45 | 117.53 | O4—C57—N14 | 129.41 |
C42—C49—C48 | 120.98 | C52—N13—N14 | 107.77 |
C42—C49—H50 | 121.46 | C52—C53—H54 | 109.95 |
C48—C49—H50 | 117.56 | C52—C53—H55 | 110.58 |
C27—C26—O3 | 127.20 | C52—C53—H56 | 110.04 |
C27—C26—N5 | 103.55 | H54—C53—H55 | 109.36 |
O3—C26—N5 | 129.25 | H54—C53—H56 | 109.65 |
C27—C28—N6 | 110.30 | H55—C53—H56 | 107.22 |
C27—C28—C29 | 126.66 | C57—N14—N13 | 111.93 |
N6—C28—C29 | 123.04 | C57—N14—C58 | 129.51 |
C43—C45—H46 | 120.86 | N13—N14—C58 | 118.56 |
C43—C45—C47 | 120.61 | N14—C58—C59 | 119.53 |
H46—C45—C47 | 118.51 | N14—C58—C67 | 120.48 |
C49—C48—C47 | 120.85 | C59—C58—C67 | 119.99 |
C49—C48—Cl2 | 119.56 | C58—C59—H60 | 119.54 |
C47—C48—Cl2 | 119.60 | C58—C59—C61 | 119.52 |
C26—N5—N6 | 112.02 | H60—C59—C61 | 120.93 |
C26—N5—C25 | 129.34 | C58—C67—C65 | 119.50 |
N6—N5—C25 | 118.64 | C58—C67—H68 | 119.77 |
C28—N6—N5 | 107.71 | C65—C67—H68 | 120.72 |
C28—C29—H30 | 110.78 | C59—C61—H62 | 118.76 |
C28—C29—H31 | 110.28 | C59—C61—C63 | 120.93 |
C28—C29—H32 | 110.29 | H62—C61—C63 | 120.30 |
H30—C29—H31 | 108.99 | C67—C65—C63 | 120.89 |
H30—C29—H32 | 109.17 | C67—C65—H66 | 118.61 |
H31—C29—H32 | 107.24 | C63—C65—H66 | 120.48 |
C45—C47—C48 | 118.33 | C61—C63—C65 | 119.16 |
C45—C47—N10 | 120.58 | C61—C63—H64 | 121.21 |
C48—C47—N10 | 121.09 | C65—C63—H64 | 119.63 |
Footnotes
‡Present address: Cosmolux International S.A., R&D / QC, rue Johan Bertels; Z.I.; 6468 Echternach, Luxembourg
1This glide plane is denoted as g, not as d, because a d glide plane with a translational vector of (a + b)/4 requires the presence of a second 'd' glide plane with a translation vector of (a − b)/4, which is absent in this model.
Acknowledgements
The authors thank David Urmann (formerly Goethe University, Frankfurt) for crystallization experiments and preliminary analyses. Harald Schweizer (Sanofi, Frankfurt) is acknowledged for single-crystal X-ray measurements of different crystals of the β and ɛ1 phases. The authors thank Christian Czech (formerly at Goethe University, Frankfurt) and Jacco van de Streek (Avant-garde Materials Simulation, Merzhausen) for the lattice-energy minimizations. Sonja Hammer (Goethe University, Frankfurt) is acknowledged for structure solution trials from powder X-ray diffraction data. Alke Meents (DESY, Hamburg) is acknowledged for attempts of extremely small single crystals using synchrotron radiation. We thank the group of Michael Karas (Goethe University, Frankfurt) for LDI-MS investigations of various printings and plastic articles containing P.O.13. We would like to thank Ute Kolb (Johannes Gutenberg University, Mainz) for providing TEM instrumentation. Nicola Ashcroft, (late) Jill Bradshaw and Amanda Berry (all IUCr) are acknowledged for many years of tremendous support while publishing with the IUCr and for providing a printed cover of Acta Crystallographica Section C for analysis and identification of P.O.13 as the red pigment in it. Open access funding enabled and organized by Projekt DEAL.
Funding information
The following funding is acknowledged: Deutsche Forschungsgemeinschaft (award No. CRC 1279).
References
Barrow, M. J. (2002). Dyes Pigments, 55, 79–89. Web of Science CSD CrossRef CAS Google Scholar
Barrow, M. J., Christie, R. M. & Badcock, T. D. (2003). Dyes Pigments, 57, 99–106. Web of Science CSD CrossRef CAS Google Scholar
Barrow, M. J., Christie, R. M., Lough, A. J., Monteith, J. E. & Standring, P. N. (2000). Dyes Pigments, 45, 153–160. Web of Science CSD CrossRef CAS Google Scholar
Bekö, S. L., Thoms, S. & Schmidt, M. U. (2014). German Patent Application, DE 102014015146 A1. Google Scholar
Bürgi, H.-B., Hostettler, M., Birkedal, H. & Schwarzenbach, D. (2005). Z. Kristallogr. 220, 1066–1075. CrossRef CAS Google Scholar
Burla, M. C., Caliandro, R., Carrozzini, B., Cascarano, G. L., Cuocci, C., Giacovazzo, C., Mallamo, M., Mazzone, A. & Polidori, G. (2015). J. Appl. Cryst. 48, 306–309. Web of Science CrossRef CAS IUCr Journals Google Scholar
Copley, R. C. B., Barnett, S. A., Karamertzanis, P. G., Harris, K. D. M., Kariuki, B. M., Xu, M., Nickels, E. A., Lancaster, R. W. & Price, S. L. (2008). Cryst. Growth Des. 8, 3474–3481. Web of Science CSD CrossRef CAS Google Scholar
Czech, C., Kalinowsky, L. & Schmidt, M. U. (2017). Acta Cryst. B73, 744–755. Web of Science CrossRef IUCr Journals Google Scholar
Ershov, A. P., Lyashenko, V. D. & Grachev, I. V. (1934). Anilinokrasochnaya Prom. 4, 306–12. CAS Google Scholar
Fischer, W. & Koch, E. (2011). International Tables for Crystallography, Vol. A, Space-group symmetry, pp. 810–811. Chester: International Union of Crystallography. Google Scholar
Gasteiger, J. & Marsili, M. (1980). Tetrahedron, 36, 3219–3228. CrossRef CAS Web of Science Google Scholar
Gemmi, M. & Lanza, A. E. (2019). Acta Cryst. B75, 495–504. Web of Science CrossRef IUCr Journals Google Scholar
Gemmi, M., Mugnaioli, E., Gorelik, T. E., Kolb, U., Palatinus, L., Boullay, P., Hovmöller, S. & Abrahams, J. P. (2019). ACS Cent. Sci. 5, 1315–1329. Web of Science CrossRef CAS PubMed Google Scholar
Gorelik, T., Schmidt, M. U., Brüning, J., Bekő, S. & Kolb, U. (2009). Cryst. Growth Des. 9, 3898–3903. Web of Science CrossRef CAS Google Scholar
Gorelik, T. E., Czech, C., Hammer, S. M. & Schmidt, M. U. (2016). CrystEngComm, 18, 529–535. Web of Science CrossRef CAS Google Scholar
Gruene, T. & Mugnaioli, E. (2021). Chem. Rev. 121, 11823–11834. Web of Science CrossRef CAS PubMed Google Scholar
Hahn, Th. (2002). International Tables for Crystallography, Vol. A, Space-Group Symmetry, 5th ed. Dordrecht: Kluwer Academic Publishers. Google Scholar
Hörnig, A., Kölle, U., Schmidt, M. U., Herberich, G. E. & Englert, U. (1993). Acta Chim. Hung. 130, 405–414. Google Scholar
Hunger, K. & Schmidt, M. U. (2018). Industrial Organic Pigments. Wiley-VCH, Weinheim, Germany. Google Scholar
Ivashevskaya, S. N., van de Streek, J., Djanhan, J. E., Brüning, J., Alig, E., Bolte, M., Schmidt, M. U., Blaschka, P., Höffken, H. W. & Erk, P. (2009). Acta Cryst. B65, 212–222. Web of Science CSD CrossRef IUCr Journals Google Scholar
Kamei, Y., Shibata, H. & Mizuguchi, J. (2011). J. Imaging Sci. Technol. 55, 30508-1–30508-5. Google Scholar
Kolb, U., Gorelik, T., Kübel, C., Otten, M. T. & Hubert, D. (2007). Ultramicroscopy, 107, 507–513. Web of Science CrossRef PubMed CAS Google Scholar
Kolb, U., Krysiak, Y. & Plana-Ruiz, S. (2019). Acta Cryst. B75, 463–474. Web of Science CSD CrossRef IUCr Journals Google Scholar
Kresse, G. & Furthmüller, J. (1996a). Comput. Mater. Sci. 6, 15–50. CrossRef CAS Web of Science Google Scholar
Kresse, G. & Furthmüller, J. (1996b). Phys. Rev. B, 54, 11169–11186. CrossRef CAS Web of Science Google Scholar
Kresse, G. & Hafner, J. (1993). Phys. Rev. B, 47, 558–561. CrossRef CAS Web of Science Google Scholar
Kresse, G. & Joubert, D. (1999). Phys. Rev. B, 59, 1758–1775. Web of Science CrossRef CAS Google Scholar
Krysiak, Y., Barton, B., Marler, B., Neder, R. B. & Kolb, U. (2018). Acta Cryst. A74, 93–101. Web of Science CrossRef IUCr Journals Google Scholar
Krysiak, Y., Marler, B., Barton, B., Plana-Ruiz, S., Gies, H., Neder, R. B. & Kolb, U. (2020). IUCrJ, 7, 522–534. Web of Science CSD CrossRef ICSD CAS PubMed IUCr Journals Google Scholar
Laska, A. L. (1910). German Patent 236856. Google Scholar
Laska, A. L. (1911). Swiss Patent CH54500, Austrian Patent 50624, French Patent FR425448, British Patent GB191101730, US Patent US1001286. Google Scholar
Lonsdale, K. (1929). Proc. R. Soc. 123A, 494–515. Google Scholar
Mayo, S. L., Olafson, B. D. & Goddard, W. A. (1990). J. Phys. Chem. 94, 8897–8909. CrossRef CAS Web of Science Google Scholar
Mugnaioli, E. & Gorelik, T. E. (2019). Acta Cryst. B75, 550–563. Web of Science CrossRef IUCr Journals Google Scholar
Mustroph, H. (1987). Z. Chem. 27, 281–289. CrossRef CAS Google Scholar
Neumann, M. A., Leusen, F. J. J. & Kendrick, J. (2008). Angew. Chem. 120, 2461–2464. CrossRef Google Scholar
Neumann, M. A. & Perrin, M.-A. (2005). J. Phys. Chem. B, 109, 15531–15541. Web of Science CrossRef PubMed CAS Google Scholar
Palatinus, L., Brázda, P., Jelínek, M., Hrdá, J., Steciuk, G. & Klementová, M. (2019). Acta Cryst. B75, 512–522. Web of Science CrossRef IUCr Journals Google Scholar
Paulus, E. F. (1982). Z. Kristallogr. 160, 235–243. CrossRef CAS Web of Science Google Scholar
Pettersen, E. F., Goddard, T. D., Huang, C. C., Couch, G. S., Greenblatt, D. M., Meng, E. C. & Ferrin, T. E. (2004). J. Comput. Chem. 25, 1605–1612. Web of Science CrossRef PubMed CAS Google Scholar
Schmidt, E. & Neder, R. B. (2017). Acta Cryst. A73, 231–237. Web of Science CrossRef IUCr Journals Google Scholar
Schmidt, M. U., Dinnebier, R. E. & Kalkhof, H. J. (2007). J. Phys. Chem. B, 111, 9722–9732. Web of Science CSD CrossRef PubMed CAS Google Scholar
Schmidt, M. U. & Glinnemann, J. (2012). Z. Kristallogr. 227, 805–818. Web of Science CrossRef CAS Google Scholar
Sheldrick, G. M. (1990). Acta Cryst. A46, 467–473. CrossRef CAS Web of Science IUCr Journals Google Scholar
Sheldrick, G. M. & Schneider, T. R. (1997). Methods Enzymol. 277, 319–343. CrossRef PubMed CAS Web of Science Google Scholar
Spek, A. L. (2015). Acta Cryst. C71, 9–18. Web of Science CrossRef IUCr Journals Google Scholar
Teteruk, J. L., Glinnemann, J., Gorelik, T. E., Linden, A. & Schmidt, M. U. (2014). Acta Cryst. B70, 296–305. Web of Science CrossRef IUCr Journals Google Scholar
Tong, H., Tan, P. & Xu, N. (2015). Sci. Rep. 5, 15378. Web of Science CrossRef PubMed Google Scholar
Warshamanage, R., Linden, A., Schmidt, M. U. & Bürgi, H.-B. (2014). Acta Cryst. B70, 283–295. Web of Science CSD CrossRef IUCr Journals Google Scholar
Weber, T. & Bürgi, H.-B. (2002). Acta Cryst. A58, 526–540. Web of Science CrossRef CAS IUCr Journals Google Scholar
Weber, T. & Simonov, A. (2012). Z. Kristallogr. 227, 238–247. Web of Science CrossRef CAS Google Scholar
Welberry, R. & Weber, T. (2016). Crystallogr. Rev. 22, 2–78. Web of Science CrossRef CAS Google Scholar
Welberry, T. R. (2004). Diffuse X-ray Scattering and Models of Disorder. IUCr Monographs on Crystallography 16, Oxford University Press, Oxford, UK. Google Scholar
Whitaker, A. (1988a). Acta Cryst. C44, 1587–1590. CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
Whitaker, A. (1988b). Acta Cryst. C44, 1767–1770. CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
Yennawar, H. P., Buchwalter, M. J., Colburn, B. K. & Silverberg, L. J. (2018). Acta Cryst. E74, 363–366. Web of Science CSD CrossRef IUCr Journals Google Scholar
Zehe, C. S., Hill, J. A., Funnell, N. P., Kreger, K., van der Zwan, K. P., Goodwin, A. L., Schmidt, H.-W. & Senker, J. (2017). Angew. Chem. Int. Ed. 56, 4432–4437. Web of Science CSD CrossRef CAS Google Scholar
This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.