weak interactions in crystals\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Studying weak inter­actions in crystals at high pressures: when hardware matters

CROSSMARK_Color_square_no_text.svg

aInstitute of Solid State Chemistry and Mechanochemistry, Siberian Branch of the Russian Academy of Sciences, Kutateladze Str. 18, Novosibirsk, 630128, Russian Federation, bNovosibirsk State University, Pirogova Str. 2, Novosibirsk, 630090, Russian Federation, and cRigaku Oxford Diffraction, Monument Park, Chalgrove, OX44 7RW, England
*Correspondence e-mail: eboldyreva@yahoo.com

Edited by H. Stoeckli-Evans, University of Neuchâtel, Switzerland (Received 20 March 2018; accepted 22 March 2018; online 17 April 2018)

The quality of structural models for 1,2,4,5-tetra­bromo­benzene (TBB), C6H2Br4, based on data collected from a single crystal in a diamond anvil cell at 0.4 GPa in situ using two different diffractometers belonging to different generations have been compared, together with the effects of applying different data-processing strategies.

1. Introduction

High-pressure data are widely used for the study of inter­molecular inter­actions in crystals. In particular, high pressure can probe inter­actions and their role in stabilizing structures and their evolution across a variety of structural transformations: anisotropic structural distortion, polymorphic transitions and chemical reactions (Katrusiak, 1991[Katrusiak, A. (1991). Cryst. Res. Technol. 26, 523-531.]; Boldyreva, 2008[Boldyreva, E. V. (2008). Acta Cryst. A64, 218-231.]; Resnati et al., 2015[Resnati, G., Boldyreva, E., Bombicz, P. & Kawano, M. (2015). IUCrJ, 2, 675-690.]; Yan et al., 2018[Yan, H., Yang, F., Pan, D., Lin, Y., Hohman, J. N., Solis-Ibarra, D., Li, F. H., Dahl, J. E. P., Carlson, R. M. K., Tkachenko, B. A., Fokin, A. A., Schreiner, P. R., Galli, G., Mao, W. L., Shen, Z.-X. & Melosh, N. A. (2018). Nature, 554, 505-510.]; Parois et al., 2010[Parois, P., Moggach, S. A., Sanchez-Benitez, J., Kamenev, K. V., Lennie, A. R., Warren, J. E., Brechin, E. K., Parsons, S. & Murrie, M. (2010). Chem. Commun. 46, 1881-1883.]). The quality of diffraction data [particularly completeness and the F2/σ(F2) ratio] is critically important for obtaining reliable information on mol­ecular conformations, inter­molecular distances and even electron charge-density distribution (Veciana et al., 2018[Veciana, J., Souto, M., Gullo, M. C., Cui, H., Casati, N., Montisci, F., Jeschke, H. O., Valentí, R., Ratera, I. & Rovira, C. (2018). Chem. Eur. J., doi 10.1002cem.201800881.]; Casati et al., 2017[Casati, N., Genoni, A., Meyer, B., Krawczuk, A. & Macchi, P. (2017). Acta Cryst. B73, 584-597.], 2016[Casati, N., Kleppe, A., Jephcoat, A. P. & Macchi, P. (2016). Nat. Commun. 7, 10901.]). Really impressive progress has been achieved over the last decade in obtaining more precise structural data from mol­ecular crystal structures of increasing complexity. The improvements are related, first of all, to a new design of diamond anvil cells (DACs) with larger opening angles (Sowa & Ahsbahs, 2006[Sowa, H. & Ahsbahs, H. (2006). J. Appl. Cryst. 39, 169-175.]; Ahsbahs, 2004[Ahsbahs, H. (2004). Z. Kristallogr. 219, 305-308.]; Boehler, 2006[Boehler, R. (2006). Rev. Sci. Instrum. 77, 115103; doi: 10.1063/1.2372734.]; Moggach et al., 2008[Moggach, S. A., Allan, D. R., Parsons, S. & Warren, J. E. (2008). J. Appl. Cryst. 41, 249-251.]). The improvements also include the use of 2D detectors instead of point detectors (Ahsbahs, 2004[Ahsbahs, H. (2004). Z. Kristallogr. 219, 305-308.]; Dubrovinsky et al., 2010[Dubrovinsky, L., Boffa-Ballaran, T., Glazyrin, K., Kurnosov, A., Frost, D., Merlini, M., Hanfland, M., Prakapenka, V. B., Schouwink, P., Pippinger, T. & Dubrovinskaia, N. (2010). High. Press. Res. 30, 620-633.]; Kantor et al., 2012[Kantor, I., Prakapenka, V., Kantor, A., Dera, P., Kurnosov, A., Sinogeikin, S., Dubrovinskaia, N. & Dubrovinsky, L. (2012). Rev. Sci. Instrum. 83, 125102.]; Dawson et al., 2004[Dawson, A., Allan, D. R., Parsons, S. & Ruf, M. (2004). J. Appl. Cryst. 37, 410-416.]), as well as applying new software for sample centering, absorption correction, recognizing and excluding unwanted reflections that do not belong to the sample, data reduction, and finding the orientation matrices for several crystallites in the same diamond anvil cell (Boldyreva et al., 2016[Boldyreva, E. V., Zakharov, B. A., Rashchenko, S. V., Seryotkin, Y. V. & Tumanov, N.A. (2016). Studying Solid-State Transformations Using In Situ X-Ray Diffraction Studies at High-Pressures. Novosibirsk: Publishing House of Siberian Branch of Russian Academy of Sciences, ISBN 978-5-7692-1526-1.]; Katrusiak, 2008[Katrusiak, A. (2008). Acta Cryst. A64, 135-148.], 2004[Katrusiak, A. (2004). Z. Kristallogr. 219, 461-467.]; Dera et al., 2013[Dera, P., Zhuravlev, K., Prakapenka, V., Rivers, M. L., Finkelstein, G. J., Grubor-Urosevic, O., Tschauner, O., Clark, S. M. & Downs, R. T. (2013). High. Press. Res. 33, 466-484.]; Casati et al., 2007[Casati, N., Macchi, P. & Sironi, A. (2007). J. Appl. Cryst. 40, 628-630.]; Angel & Gonzalez-Platas, 2013[Angel, R. & Gonzalez-Platas, J. (2013). J. Appl. Cryst. 46, 252-254.]). Special methods of data processing, in addition to precise experiments, now even make it possible to obtain data for charge-density studies (Veciana et al., 2018[Veciana, J., Souto, M., Gullo, M. C., Cui, H., Casati, N., Montisci, F., Jeschke, H. O., Valentí, R., Ratera, I. & Rovira, C. (2018). Chem. Eur. J., doi 10.1002cem.201800881.]; Casati et al., 2017[Casati, N., Genoni, A., Meyer, B., Krawczuk, A. & Macchi, P. (2017). Acta Cryst. B73, 584-597.], 2016[Casati, N., Kleppe, A., Jephcoat, A. P. & Macchi, P. (2016). Nat. Commun. 7, 10901.]), and to follow related changes with pressure. This has been demonstrated for example by following the reduction in aromaticity of syn-1,6:8,13-bis­carbon­yl[14]annulene on compression (Casati et al., 2016[Casati, N., Kleppe, A., Jephcoat, A. P. & Macchi, P. (2016). Nat. Commun. 7, 10901.]). Advances in the quality of high-pressure data for mol­ecular crystals have often been related to the use of synchrotron radiation. However, with limited access to synchrotrons, in-house experiments remain the most common type of high-pressure experiments for organic solids.

A new generation of laboratory diffractometers has been developed recently that makes it possible to collect data at high pressures from even small and weakly diffracting crystals. In this contribution, we present the results of a comparison of the data collected using two different diffractometers that were manufactured by the same company within a 10 year inter­val (Fig. 1[link]). The first is an XtaLAB Synergy-S Dualflex diffractometer with Ag Kα radiation (PhotonJet-S source) and Pilatus3 X CdTe 300K hybrid photon-counting (HPC) detector from Dectris that was manufactured by Rigaku Oxford Diffraction in 2017, while the second is an Oxford Diffraction Gemini R Ultra diffractometer with Mo Kα radiation (Enhance X-ray source) and Ruby charge-coupled device (CCD) detector, manufactured by Oxford Diffraction in 2007. The main parameters characterizing the two instruments are compared in Table 1[link]. We have collected data on the two different instruments from the same sample at the same pressure in the same DAC. We have also compared the results of applying different strategies for the data reduction.

Table 1
Comparison of technical characteristics of the diffractometers

  XtaLAB Synergy-S Dualflex Oxford Diffraction Gemini R Ultra
Radiation type Ag Kα Mo Kα
X-ray source type PhotonJet-S source Enhance X-ray source
Beam characteristics 0.12 mm beam 0.5 mm beam
X-ray optics double-bounce multilayer optics graphite monochromator
Detector model Pilatus3 X CdTe 300K Ruby
Detector type HPC – photon counting CCD – integrative detector
Quantum efficiency >90% >80%
Read-out frequency (Hz) 20 <0.3
Goniometer four-circle Kappa goniometer (new generation) four-circle Kappa goniometer
Data collection mode shutterless data collection shuttered data collection
[Figure 1]
Figure 1
Cabinet view of diffactometers used: (a) XtaLAB Synergy-S Dualflex; (b) Oxford Diffraction Gemini R Ultra.

As a sample we selected single crystals of 1,2,4,5-tetra­bromo­benzene (TBB). TBB is a well-known thermosalient compound, which exhibits large, spontaneous mechanical response across the phase transition on heating (Sahoo et al., 2013[Sahoo, S. C., Sinha, S. B., Kiran, M. S. R. N., Ramamurty, U., Dericioglu, A. F., Reddy, C. M. & Naumov, P. (2013). J. Am. Chem. Soc. 135, 13843-13850.]; Zakharov et al., 2018[Zakharov, B. A., Michalchuk, A. A. L., Morrison, C. A. & Boldyreva, E. V. (2018). Phys. Chem. Chem. Phys., doi: 10.1039/C7CP08609A.] and references therein). It has been shown recently that data on the high-pressure behaviour of such materials can be helpful in order to understand the origin of the thermosalient effect (Zakharov et al., 2017[Zakharov, B. A., Gribov, P. A., Matvienko, A. A. & Boldyreva, E. V. (2017). Z. Kristallogr. 232, 751-757.]). TBB crystallizes in the monoclinic space group P21/n. Being a thermosalient material, it shows a significant mechanical response, even though the phase transition on heating is accompanied by only minute rearrangements at the mol­ecular level and only minimal changes in the inter­molecular contacts (Sahoo et al., 2013[Sahoo, S. C., Sinha, S. B., Kiran, M. S. R. N., Ramamurty, U., Dericioglu, A. F., Reddy, C. M. & Naumov, P. (2013). J. Am. Chem. Soc. 135, 13843-13850.]; Zakharov et al., 2018[Zakharov, B. A., Michalchuk, A. A. L., Morrison, C. A. & Boldyreva, E. V. (2018). Phys. Chem. Chem. Phys., doi: 10.1039/C7CP08609A.]). This makes it important to have high-quality structural data at multiple pressure and temperature (PT) conditions when studying the role of the inter­molecular inter­actions in the thermosalient effect. High noise level, low data completeness, low F2/σ(F2) and data-to-number of parameters ratios can lead to the loss of most of the information related to the electron-density distribution in the crystal. When using `older-generation' in-house diffractometers, low data quality can make it impossible to refine the crystal structure in even an isotropic approximation. Therefore, fine details in the orientation of anisotropic displacement parameters (ADPs) and precise values for the inter­atomic distance changes, which are of great importance for studying the mechanical response of the crystal to variations in PT conditions, will not be accessible. The new-generation instruments are expected to improve the quality of the diffraction data and the structural models based on the refinement of these data. At the same time, using a newer instrument alone does not guarantee a high-quality structural model. The data-processing strategy is critically important for data collected from a sample in a DAC at high pressure (Boldyreva et al., 2016[Boldyreva, E. V., Zakharov, B. A., Rashchenko, S. V., Seryotkin, Y. V. & Tumanov, N.A. (2016). Studying Solid-State Transformations Using In Situ X-Ray Diffraction Studies at High-Pressures. Novosibirsk: Publishing House of Siberian Branch of Russian Academy of Sciences, ISBN 978-5-7692-1526-1.]; Katrusiak, 2008[Katrusiak, A. (2008). Acta Cryst. A64, 135-148.], 2004[Katrusiak, A. (2004). Z. Kristallogr. 219, 461-467.]; Dera et al., 2013[Dera, P., Zhuravlev, K., Prakapenka, V., Rivers, M. L., Finkelstein, G. J., Grubor-Urosevic, O., Tschauner, O., Clark, S. M. & Downs, R. T. (2013). High. Press. Res. 33, 466-484.]; Casati et al., 2007[Casati, N., Macchi, P. & Sironi, A. (2007). J. Appl. Cryst. 40, 628-630.]; Angel & Gonzalez-Platas, 2013[Angel, R. & Gonzalez-Platas, J. (2013). J. Appl. Cryst. 46, 252-254.]). These data are inevitably `contaminated' by absorption of X-rays by the materials of the DAC (diamond, metal) and reflections originating from diffraction of the diamonds, gasket or the ruby calibrant. The presence of these reflections also corrupts the measured intensities of the sample reflections, either by direct overlap or because they may have an influence on the estimated background level. The aim of this study was to compare the data quality collected from the same sample in a DAC at high pressure in situ using diffractometers belonging to different generations. For data collected using both of the two instruments, we have used several different strategies for the data processing. The aim of this was to test the relative importance of applying different techniques for correction of the raw data for increasing the reliability and improving the quality of the structural model.

2. Experimental

Single crystals of 1,2,4,5-tetra­bromo­benzene (TBB) were prepared by slow evaporation of chloro­form solutions, using 200 mg of TBB (Sigma–Aldrich, 97%) dissolved in 9 ml of chloro­form at room temperature.

[Scheme 1]

The sample was mounted in an Almax Boehler DAC (Boehler, 2006[Boehler, R. (2006). Rev. Sci. Instrum. 77, 115103; doi: 10.1063/1.2372734.]). A stainless steel sheet with an initial thickness of 200 µm was pre-indented to 100 µm and used as a gasket. The ruby fluorescence method was used for pressure calibration (Forman et al., 1972[Forman, R. A., Piermarini, G. J., Dean Barnett, J. & Block, S. (1972). Science, 176, 284-285.]; Piermarini et al., 1975[Piermarini, G. J., Block, S., Barnett, J. D. & Forman, R. A. (1975). J. Appl. Phys. 46, 2774-2780.]). A methanol–ethanol mixture (4:1) was used as hydro­static pressure-transmitting medium (Piermarini et al., 1973[Piermarini, G. J., Block, S. & Barnett, J. D. (1973). J. Appl. Phys. 44, 5377-5382.]; Angel et al., 2007[Angel, R. J., Bujak, M., Zhao, J., Gatta, G. D. & Jacobsen, S. D. (2007). J. Appl. Cryst. 40, 26-32.]).

Single-crystal X-ray diffraction data were collected on the same crystal in the same DAC at a hydro­static pressure of 0.4 GPa. Data were collected using two different instruments: (1) an XtaLAB Synergy-S Dualflex diffractometer with Ag Kα radiation (PhotonJet-S source) and Pilatus3 X CdTe 300K HPC detector from Dectris (manufactured by Rigaku Oxford Diffraction in 2017), and (2) an Oxford Diffraction Gemini R Ultra diffractometer with Mo Kα radiation (Enhance X-ray source) and Ruby CCD detector (manufactured by Oxford Diffraction in 2007). Data collection, cell refinement and data reduction were performed using CrysAlis PRO software (Rigaku OD, 2016[Rigaku OD (2016). CrysAlis PRO. Rigaku Oxford Diffraction Ltd, Yarnton, England.]). Multiple strategies were tried on each instrument. Some of the strategies deliberately neglected good-practice techniques of introducing certain high-pressure data corrections in order to evaluate the extent to which this neglect can worsen the data quality.

For data collection (1), X-ray diffraction data were treated and attempts were made to refine the structure in three different ways:

(a) Gaussian absorption correction using ABSORB-7 (Angel & Gonzalez-Platas, 2013[Angel, R. & Gonzalez-Platas, J. (2013). J. Appl. Cryst. 46, 252-254.]) implemented in CrysAlis PRO software (Rigaku OD, 2016[Rigaku OD (2016). CrysAlis PRO. Rigaku Oxford Diffraction Ltd, Yarnton, England.]). Both crystal and DAC absorption were taken into account. The most disagreeable reflections from the sample that overlapped with diamond and gasket reflections were not excluded from the HKL file. All non-H atoms were refined anisotropically.

(b) Gaussian absorption correction using ABSORB-7 (Angel & Gonzalez-Platas, 2013[Angel, R. & Gonzalez-Platas, J. (2013). J. Appl. Cryst. 46, 252-254.]) implemented in CrysAlis PRO software (Rigaku OD, 2016[Rigaku OD (2016). CrysAlis PRO. Rigaku Oxford Diffraction Ltd, Yarnton, England.]). Both crystal and DAC absorption were taken into account. The most disagreeable reflections from the sample that overlapped with diamond and gasket reflections were excluded manually from the HKL file. All non-H atoms were refined anisotropically.

(c) Spherical absorption correction as implemented in CrysAlis PRO software (Rigaku OD, 2016[Rigaku OD (2016). CrysAlis PRO. Rigaku Oxford Diffraction Ltd, Yarnton, England.]). Only crystal absorption was taken into account. The most disagreeable reflections from the sample that overlapped with diamond and gasket reflections were manually excluded from the HKL file. All non-H atoms were refined anisotropically.

For data collection (2), X-ray diffraction data were treated and attempts were made to refine in six different ways:

(d) the same as for (a).

(e) the same as for (b).

(f) the same as for (c).

(g) the same as for (a), but carbon atoms were refined isotropically.

(h) the same as for (b), but carbon atoms were refined isotropically.

(i) the same as for (c), but carbon atoms were refined isotropically.

For all the refinements at high pressure, the initial crystal structure model was taken from single-crystal diffraction data at ambient conditions (Zakharov et al., 2018[Zakharov, B. A., Michalchuk, A. A. L., Morrison, C. A. & Boldyreva, E. V. (2018). Phys. Chem. Chem. Phys., doi: 10.1039/C7CP08609A.]). Refinements were carried out with SHELXL2018/1 (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]) using Shelxle (Hübschle et al., 2011[Hübschle, C. B., Sheldrick, G. M. & Dittrich, B. (2011). J. Appl. Cryst. 44, 1281-1284.]) as the GUI without any restraints. Hydrogen-atom parameters were constrained using AFIX 43 with Uiso(H) = 1.2Ueq(C). Mercury (Macrae et al., 2008[Macrae, C. F., Bruno, I. J., Chisholm, J. A., Edgington, P. R., McCabe, P., Pidcock, E., Rodriguez-Monge, L., Taylor, R., van de Streek, J. & Wood, P. A. (2008). J. Appl. Cryst. 41, 466-470.]), checkCIF/PLATON (Spek, 2009[Spek, A. L. (2009). Acta Cryst. D65, 148-155.]) and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]) were used for structure visualization, analysis and preparation of the CIF files for publication.

3. Results and discussion

Crystal data, data collection and refinement parameters are summarized in Table 2[link]. In comparison with the older Gemini R Ultra device, used for data collection (2), the Synergy-S diffractometer, used for data collection (1), was superior for data collection. Compared to instrument (2), collection of single-crystal X-ray data on (1) was much faster (6 vs 32 h), with a higher F2/σ(F2) ratio (18 vs. 10) and data completeness (68 vs 58%). A higher HKL range allowed us to increase the number of reflections used for cell-parameter refinement by a factor of 1.5. The resulting values of the lattice parameters appear to be almost the same in the two cases: the largest difference, 0.2%, was observed for lattice parameter b. Standard uncertainties for the cell parameters were slightly higher for (1) than for (2). This is presumably related to the smaller 2θ values for stronger reflections owing to the use of the harder Ag Kα radiation.

Table 2
Experimental details

For all structures: C6H2Br4, Mr = 393.72, monoclinic, P21/n, Z = 2. Experiments were carried out at 293 K. Crystal size 0.18 × 0.07 × 0.01 (mm). H-atom parameters were constrained. Refinements not acceptable for publication (incorrect) are highlighted in red, preferable in green, and those publishable but not always preferable are not highlighted.

  (a) Ag, ABSORB-7, raw (b) Ag, ABSORB-7 (c) Ag, CA sphere
Crystal data
a, b, c (Å) 3.9390 (9), 10.781 (4), 9.944 (4) 3.9390 (9), 10.781 (4), 9.944 (4) 3.9390 (9), 10.781 (4), 9.944 (4)
β (°) 100.49 (3) 100.49 (3) 100.49 (3)
V3) 415.2 (2) 415.2 (2) 415.2 (2)
Radiation type Ag Kα, λ = 0.56087 Å Ag Kα, λ = 0.56087 Å Ag Kα, λ = 0.56087 Å
No. of reflections for cell measurement 748 748 748
θ range (°) for cell measurement 2.2–22.9 2.2–22.9 2.2–22.9
μ (mm−1) 10.33 10.33 10.33
 
Data collection [total experiment time = 6 hours, exposure time = 45 seconds, F2/σ(F2) = 18, data completeness = 68% (inf = 0.8 Å)]
Absorption correction Gaussian Gaussian Sphere
Tmin, Tmax 0.486, 0.562 0.486, 0.562 0.638, 0.645
No. of measured, independent and observed [I > 2σ(I)] reflections 2503, 893, 513 2445, 870, 496 2453, 870, 494
Rint 0.048 0.048 0.050
(sin θ/λ)max−1) 0.801 0.801 0.801
Range of h, k, l h = −5→6, k = −14→14, l = −11→12 h = −5→6, k = −14→14, l = −11→12 h = −5→6, k = −14→14, l = −11→12
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.047, 0.206, 1.02 0.037, 0.073, 0.93 0.037, 0.071, 0.91
No. of reflections 893 870 870
No. of parameters 46 46 46
(Δ/σ)max 0.014 0.001 < 0.001
Δρmax, Δρmin (e Å−3) 1.55, −1.48 0.54, −0.54 0.53, −0.49
  (d) Mo, ABSORB-7, raw (e) Mo, ABSORB-7 (f) Mo, CA sphere
Crystal data
a, b, c (Å) 3.9431 (5), 10.7566 (18), 9.964 (2) 3.9431 (5), 10.7566 (18), 9.964 (2) 3.9431 (5), 10.7566 (18), 9.964 (2)
β (°) 100.557 (15) 100.557 (15) 100.557 (15)
V3) 415.47 (13) 415.47 (13) 415.47 (13)
Radiation type Mo Kα Mo Kα Mo Kα
No. of reflections for cell measurement 514 514 514
θ range (°) for cell measurement 2.8–22.4 2.8–22.4 2.8–22.4
μ (mm−1) 19.29 19.29 19.29
 
Data collection [total experiment time = 32 hours, exposure time = 60 seconds, F2/σ(F2) = 10, data completeness = 58% (inf = 0.8 Å)]
Absorption correction Gaussian Gaussian Sphere
Tmin, Tmax 0.361, 0.434 0.361, 0.434 0.638, 0.645
No. of measured, independent and observed [I > 2σ(I)] reflections 2177, 550, 323 2116, 531, 313 2125, 531, 319
Rint 0.105 0.103 0.102
(sin θ/λ)max−1) 0.663 0.663 0.663
Range of h, k, l h = −5→5, k = −12→11, l = −10→10 h = −5→5, k = −12→11, l = −10→10 h = −5→5, k = −12→11, l = −10→10
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.101, 0.347, 1.19 0.071, 0.169, 1.05 0.069, 0.157, 1.05
No. of reflections 550 531 531
No. of parameters 46 46 46
(Δ/σ)max 0.089 0.592 0.523
Δρmax, Δρmin (e Å−3) 2.65, −2.89 1.04, −0.89 0.93, −0.83
  (g) Mo, ABSORB-7, raw, C iso (h) Mo, ABSORB-7, C iso (i) Mo, CA sphere, C iso
Crystal data
a, b, c (Å) 3.9431 (5), 10.7566 (18), 9.964 (2) 3.9431 (5), 10.7566 (18), 9.964 (2) 3.9431 (5), 10.7566 (18), 9.964 (2)
β (°) 100.557 (15) 100.557 (15) 100.557 (15)
V3) 415.47 (13) 415.47 (13) 415.47 (13)
Radiation type Mo Kα Mo Kα Mo Kα
No. of reflections for cell measurement 514 514 514
θ range (°) for cell measurement 2.8–22.4 2.8–22.4 2.8–22.4
μ (mm−1) 19.29 19.29 19.29
 
Data collection [total experiment time = 32 hours, exposure time = 60 seconds, F2/σ(F2) = 10, data completeness = 58% (inf = 0.8 Å)]
Absorption correction Gaussian Gaussian Sphere
Tmin, Tmax 0.361, 0.434 0.361, 0.434 0.638, 0.645
No. of measured, independent and observed [I > 2σ(I)] reflections 2177, 550, 323 2116, 531, 313 2125, 531, 319
Rint 0.105 0.103 0.102
(sin θ/λ)max−1) 0.663 0.663 0.663
Range of h, k, l h = −5→5, k = −12→11, l = −10→10 h = −5→5, k = −12→11, l = −10→10 h = −5→5, k = −12→11, l = −10→10
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.097, 0.345, 1.17 0.073, 0.177, 1.04 0.071, 0.167, 1.03
No. of reflections 550 531 531
No. of parameters 31 31 31
(Δ/σ)max < 0.001 < 0.001 < 0.001
Δρmax, Δρmin (e Å−3) 2.65, −2.90 1.03, −0.88 0.95, −0.82

Shorter wavelengths are generally prefered for samples mounted in a DAC with a fixed window-opening size. From a data completeness point of view, this provides the same number of reflections in a narrower 2θ range. Ag Kα radiation is therefore becoming popular for high-pressure X-ray diffraction studies (Saouane et al., 2013[Saouane, S., Norman, S. E., Hardacre, C. & Fabbiani, F. P. A. (2013). Chem. Sci. 4, 1270-1280.]; Saouane & Fabbiani, 2015[Saouane, S. & Fabbiani, F. P. A. (2015). Cryst. Growth Des. 15, 3875-3884.]; Granero-García et al., 2017[Granero-García, R., Falenty, A. & Fabbiani, F. P. A. (2017). Chem. Eur. J. 23, 3691-3698.]). The number of independent reflections for data collection (1) was 1.6 times greater than for (2) (893 vs 550), as a result of using a shorter wavelength. The more efficient HPC detector and the brighter X-ray source allowed us to measure reflection intensities with higher precision. This gave us a twofold lower Rint value for data collection (1): 0.048 for data set (b) vs 0.105 for data sets (e) and (h).

Displacement ellipsoid plots for the different methods of data treatment and refinement are shown in Fig. 2[link]. Taking into account the refinement data presented in Table 2[link], one can conclude that the best results are provided by refinements (b) and (c), where the use of a modern device permitted a more precise and faster measurement of the intensities of the diffraction reflections. The quality of the diffraction data enabled a crystal-structure refinement in the anisotropic approximation for all non-H atoms, providing reasonable values and shapes of the displacement ellipsoids. For the refinement variant (a), for which the sample reflections that overlapped with diamond and gasket reflections were not excluded from the HKL file, the refinement did not converge, and when an anisotropic refinement was attempted a non-positive-definite atomic displacement ellipsoid was obtained for one of the carbon atoms.

[Figure 2]
Figure 2
Displacement ellipsoid plots for 1,2,4,5-tetra­bromo­benzene mol­ecules obtained with different data-treatment procedures. Carbon atoms for structure refinements (g), (h) and (i) were refined using an isotropic approximation. Cubes show atoms with negative thermal parameters. Refinements marked V are preferable for publication; those marked W are publishable but not always preferable, and those marked X are not acceptable for publication (incorrect).

For data collection (2), the refinement results were of much lower quality than those for data collection (1). As expected, the worst results were provided by refinements (d) and (g) for which the sample reflections that overlapped with the diamond and gasket reflections were not excluded from the HKL file. The refinement did not converge, and two of the carbon atoms were characterized by non-positive-definite ellipsoids when attempting to use an anisotropic model. Removal of the corrupted reflections from the HKL file did not improve refinement results. The anisotropic thermal parameters were still not adequate for the (e) and (f) refinements. Publishable refinement results in this case of impossible anisotropic refinement could be obtained in two ways: viz. by applying SHELX restraints for the thermal parameters of carbon atoms, e.g. SIMU and DELU, with low standard uncertainty values, or by refining the carbon atoms in an isotropic approximation, as was done for the (h) and (i) refinements.

Different absorption correction types were tested for both data collection strategies. The refinement results provided by the Gaussian and spherical absorption corrections are defined as (b) and (c), (e) and (f), (h) and (i), respectively. One can see that the R-factors are comparable and acceptable for both absorption-correction strategies. A potential explanation for the similarity of the Gaussian and spherical absorption correction results for data collection (1) rests in the fact that TBB is a medium-absorbing sample (μ is 10.33 mm−1 for Ag Kα). In the case of data collection (2), TBB is much more absorbing (μ is 19.29 mm−1 for Mo Kα radiation) but the overall data quality is low (intensities are not measured precisely) and even the good-practice procedure of applying an absorption correction does not improve data quality. Generally, it is preferable to use a Gaussian absorption correction (both for the crystal and for the DAC), especially for strongly absorbing samples since it calculates the `true' transmission factors using the actual crystal and DAC geometries. For example, data sets (b) and (h), and (e) in the case of reasonable anisotropic thermal displacement parameters, would be the most preferable for the experimental set-up described.

4. Conclusions

In order to obtain reliable information on inter­molecular inter­actions in a crystal structure, one needs high-quality data. This is especially critical for data collected in a DAC at high pressure, when data completeness and the availability of reciprocal space are limited. A comparison of the results obtained using different instruments and different data-processing methods has illustrated that the data processing itself plays a crucial role in obtaining reliable results. At the same time, a modern instrument belonging to the new generation makes it possible to speed up data collection, increase the signal-to-noise intensity ratio and the number of observed reflections, and with shorter wavelength data completeness for a sample mounted in a DAC. Data collection for the 1,2,4,5-tetra­bromo­benzene crystal mounted in a DAC using a modern XtaLAB Synergy-S Dualflex diffractometer with Ag Kα radiation and a Pilatus3 X CdTe 300K HPC detector took six hours, and allowed us to obtain high-quality data for an anisotropic crystal-structure refinement without any restraints.

Using the older diffractometer from the previous generation, an Oxford Diffraction Gemini R Ultra with Mo Kα radiation and a Ruby CCD detector, did not allow us to obtain diffraction data of the same quality, even when using a higher exposure time, for which data collection took 32 h; the anisotropic refinement was possible only for the heavier bromine atoms. The carbon atoms could be refined reasonably only in an isotropic approximation, or by restraining their thermal parameters. Data completeness, HKL ranges and the F2/σ(F2) ratio were lower, and the R-factors were higher compared to the values obtained when using the modern XtaLAB Synergy-S Dualflex diffractometer described above.

Crystal-structure refinement using the same primary data set, but different data-reduction strategies has revealed that eliminating the sample reflections with wrong intensities (affected by the presence of diamond, as well as powder-diffraction rings originating from the metal gasket) is the most important correction of primary data. The exact procedure for the absorption correction was less critical in the particular case considered in this work. However, generally and especially for strong absorbers, a Gaussian absorption correction both for the crystal and the DAC data can help to increase the quality of the refinement significantly, since it calculates the `true' transmission factors using the actual crystal and DAC geometries.

Supporting information


Computing details top

For all structures, data collection: CrysAlis PRO (Rigaku OD, 2016); cell refinement: CrysAlis PRO (Rigaku OD, 2016); data reduction: CrysAlis PRO (Rigaku OD, 2016). Program(s) used to refine structure: SHELXL2018 (Sheldrick, 2015) for Ag-Absorb7-raw_a; SHELXL2018/1 (Sheldrick, 2015) for Ag-Absorb7_b, Ag-CAsphere_c, Mo-Absorb7-raw_d, Mo-Absorb7_e, Mo-CAsphere_f, Mo-Absorb7-raw-Ciso_g, Mo-Absorb7-Ciso_h, Mo-CAsphere-Ciso_i. For all structures, molecular graphics: Mercury (Macrae et al., 2008). Software used to prepare material for publication: SHELXL2018 (Sheldrick, 2015) and publCIF (Westrip, 2010) for Ag-Absorb7-raw_a; SHELXL2018/1 (Sheldrick, 2015) and publCIF (Westrip, 2010) for Ag-Absorb7_b, Ag-CAsphere_c, Mo-Absorb7-raw_d, Mo-Absorb7_e, Mo-CAsphere_f, Mo-Absorb7-raw-Ciso_g, Mo-Absorb7-Ciso_h, Mo-CAsphere-Ciso_i.

1,2,4,5-Tetrabromobenzene (Ag-Absorb7-raw_a) top
Crystal data top
C6H2Br4F(000) = 356
Mr = 393.72Dx = 3.149 Mg m3
Monoclinic, P21/nAg Kα radiation, λ = 0.56087 Å
a = 3.9390 (9) ÅCell parameters from 748 reflections
b = 10.781 (4) Åθ = 2.2–22.9°
c = 9.944 (4) ŵ = 10.33 mm1
β = 100.49 (3)°T = 293 K
V = 415.2 (2) Å3Block, colourless
Z = 20.18 × 0.07 × 0.01 mm
Data collection top
XtaLAB Synergy, Dualflex, Pilatus 300K
diffractometer
513 reflections with I > 2σ(I)
ω–scanRint = 0.048
Absorption correction: gaussian
[CrysAlis PRO (Rigaku OD, 2016) and ABSORB (Angel et al., 2007)]
θmax = 26.7°, θmin = 2.2°
Tmin = 0.486, Tmax = 0.562h = 56
2503 measured reflectionsk = 1414
893 independent reflectionsl = 1112
Refinement top
Refinement on F20 restraints
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.047H-atom parameters constrained
wR(F2) = 0.206 w = 1/[σ2(Fo2) + (0.1247P)2]
where P = (Fo2 + 2Fc2)/3
S = 1.02(Δ/σ)max = 0.014
893 reflectionsΔρmax = 1.55 e Å3
46 parametersΔρmin = 1.48 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Br10.5830 (3)0.79104 (12)0.59419 (15)0.0407 (5)
Br20.3468 (3)0.57488 (12)0.80180 (13)0.0377 (5)
C10.533 (3)0.6214 (13)0.5426 (15)0.035 (3)
C20.441 (3)0.5314 (12)0.6295 (13)0.029 (3)
C30.402 (3)0.4136 (10)0.5879 (15)0.034 (3)
H30.3317940.3541650.6448300.041*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Br10.0582 (8)0.0262 (10)0.0386 (12)0.0032 (5)0.0110 (6)0.0034 (4)
Br20.0498 (7)0.0380 (11)0.0280 (11)0.0009 (5)0.0142 (6)0.0026 (4)
C10.023 (5)0.040 (10)0.038 (10)0.008 (5)0.001 (5)0.009 (5)
C20.028 (5)0.041 (10)0.021 (10)0.001 (5)0.011 (5)0.008 (4)
C30.023 (5)0.002 (8)0.079 (12)0.008 (4)0.014 (5)0.001 (4)
Geometric parameters (Å, º) top
Br1—C11.900 (14)C1—C3i1.419 (19)
Br2—C21.878 (12)C2—C31.336 (16)
C1—C21.389 (19)C3—H30.9300
C2—C1—C3i119.5 (12)C1—C2—Br2120.7 (10)
C2—C1—Br1122.0 (11)C2—C3—C1i120.4 (11)
C3i—C1—Br1118.4 (10)C2—C3—H3119.8
C3—C2—C1120.0 (13)C1i—C3—H3119.8
C3—C2—Br2119.3 (10)
C3i—C1—C2—C32.6 (19)Br1—C1—C2—Br22.2 (14)
Br1—C1—C2—C3178.7 (8)C1—C2—C3—C1i2.7 (19)
C3i—C1—C2—Br2179.1 (9)Br2—C2—C3—C1i179.2 (9)
Symmetry code: (i) x+1, y+1, z+1.
1,2,4,5-tetrabromobenzene (Ag-Absorb7_b) top
Crystal data top
C6H2Br4F(000) = 356
Mr = 393.72Dx = 3.149 Mg m3
Monoclinic, P21/nAg Kα radiation, λ = 0.56087 Å
a = 3.9390 (9) ÅCell parameters from 748 reflections
b = 10.781 (4) Åθ = 2.2–22.9°
c = 9.944 (4) ŵ = 10.33 mm1
β = 100.49 (3)°T = 293 K
V = 415.2 (2) Å3Block, colourless
Z = 20.18 × 0.07 × 0.01 mm
Data collection top
XtaLAB Synergy, Dualflex, Pilatus 300K
diffractometer
496 reflections with I > 2σ(I)
ω–scanRint = 0.048
Absorption correction: gaussian
(CrysAlisPro; Rigaku OD, 2016) and (ABSORB; Angel et al., 2007)
θmax = 26.7°, θmin = 2.2°
Tmin = 0.486, Tmax = 0.562h = 56
2445 measured reflectionsk = 1414
870 independent reflectionsl = 1112
Refinement top
Refinement on F20 restraints
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.037H-atom parameters constrained
wR(F2) = 0.073 w = 1/[σ2(Fo2) + (0.023P)2]
where P = (Fo2 + 2Fc2)/3
S = 0.93(Δ/σ)max = 0.001
870 reflectionsΔρmax = 0.54 e Å3
46 parametersΔρmin = 0.54 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Br10.58340 (14)0.79093 (6)0.59432 (7)0.0409 (2)
Br20.34699 (14)0.57466 (6)0.80194 (7)0.0380 (2)
C10.5332 (12)0.6218 (6)0.5414 (7)0.0314 (15)
C20.4386 (12)0.5325 (5)0.6290 (6)0.0276 (14)
C30.4002 (12)0.4113 (5)0.5857 (7)0.0297 (14)
H30.3293280.3513720.6419330.036*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Br10.0583 (4)0.0255 (5)0.0399 (6)0.0036 (3)0.0115 (3)0.0034 (2)
Br20.0502 (3)0.0372 (5)0.0291 (5)0.0015 (3)0.0139 (3)0.0027 (2)
C10.026 (2)0.020 (5)0.046 (6)0.003 (2)0.001 (3)0.004 (2)
C20.027 (3)0.036 (5)0.022 (6)0.004 (3)0.010 (3)0.002 (2)
C30.033 (3)0.017 (5)0.041 (6)0.004 (3)0.013 (3)0.005 (2)
Geometric parameters (Å, º) top
Br1—C11.899 (6)C1—C21.394 (7)
Br2—C21.876 (6)C2—C31.377 (8)
C1—C3i1.382 (8)C3—H30.9300
C3i—C1—C2120.7 (6)C1—C2—Br2121.6 (5)
C3i—C1—Br1118.2 (4)C2—C3—C1i120.2 (5)
C2—C1—Br1121.0 (5)C2—C3—H3119.9
C3—C2—C1119.1 (6)C1i—C3—H3119.9
C3—C2—Br2119.2 (4)
C3i—C1—C2—C32.3 (9)Br1—C1—C2—Br21.3 (6)
Br1—C1—C2—C3179.1 (4)C1—C2—C3—C1i2.2 (9)
C3i—C1—C2—Br2180.0 (4)Br2—C2—C3—C1i179.9 (4)
Symmetry code: (i) x+1, y+1, z+1.
1,2,4,5-tetrabromobenzene (Ag-CAsphere_c) top
Crystal data top
C6H2Br4F(000) = 356
Mr = 393.72Dx = 3.149 Mg m3
Monoclinic, P21/nAg Kα radiation, λ = 0.56087 Å
a = 3.9390 (9) ÅCell parameters from 748 reflections
b = 10.781 (4) Åθ = 2.2–22.9°
c = 9.944 (4) ŵ = 10.33 mm1
β = 100.49 (3)°T = 293 K
V = 415.2 (2) Å3Block, colourless
Z = 20.18 × 0.07 × 0.01 mm
Data collection top
XtaLAB Synergy, Dualflex, Pilatus 300K
diffractometer
494 reflections with I > 2σ(I)
ω–scanRint = 0.050
Absorption correction: for a sphere
(CrysAlisPro; Rigaku OD, 2016)
θmax = 26.7°, θmin = 2.2°
Tmin = 0.638, Tmax = 0.645h = 56
2453 measured reflectionsk = 1414
870 independent reflectionsl = 1112
Refinement top
Refinement on F20 restraints
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.037H-atom parameters constrained
wR(F2) = 0.071 w = 1/[σ2(Fo2) + (0.023P)2]
where P = (Fo2 + 2Fc2)/3
S = 0.91(Δ/σ)max < 0.001
870 reflectionsΔρmax = 0.53 e Å3
46 parametersΔρmin = 0.49 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Br10.58346 (14)0.79099 (6)0.59430 (7)0.0412 (2)
Br20.34702 (14)0.57465 (6)0.80195 (7)0.0383 (2)
C10.5329 (12)0.6222 (5)0.5413 (7)0.0315 (14)
C20.4384 (12)0.5328 (5)0.6291 (6)0.0284 (14)
C30.4001 (12)0.4113 (5)0.5855 (7)0.0305 (14)
H30.3287800.3515310.6417710.037*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Br10.0587 (4)0.0267 (5)0.0392 (6)0.0037 (3)0.0116 (3)0.0034 (2)
Br20.0506 (3)0.0382 (5)0.0286 (5)0.0014 (3)0.0140 (3)0.0026 (2)
C10.026 (2)0.022 (5)0.045 (6)0.003 (2)0.001 (3)0.004 (2)
C20.028 (3)0.037 (5)0.022 (6)0.004 (3)0.009 (3)0.002 (2)
C30.033 (3)0.020 (5)0.041 (6)0.004 (3)0.014 (3)0.005 (2)
Geometric parameters (Å, º) top
Br1—C11.895 (6)C1—C21.395 (7)
Br2—C21.875 (6)C2—C31.379 (7)
C1—C3i1.382 (8)C3—H30.9300
C3i—C1—C2120.5 (6)C1—C2—Br2121.7 (5)
C3i—C1—Br1118.5 (4)C2—C3—C1i120.4 (5)
C2—C1—Br1121.1 (5)C2—C3—H3119.8
C3—C2—C1119.0 (6)C1i—C3—H3119.8
C3—C2—Br2119.3 (4)
C3i—C1—C2—C32.4 (9)Br1—C1—C2—Br21.4 (6)
Br1—C1—C2—C3179.1 (4)C1—C2—C3—C1i2.4 (9)
C3i—C1—C2—Br2179.8 (4)Br2—C2—C3—C1i179.7 (4)
Symmetry code: (i) x+1, y+1, z+1.
1,2,4,5-tetrabromobenzene (Mo-Absorb7-raw_d) top
Crystal data top
C6H2Br4F(000) = 356
Mr = 393.72Dx = 3.147 Mg m3
Monoclinic, P21/nMo Kα radiation, λ = 0.71073 Å
a = 3.9431 (5) ÅCell parameters from 514 reflections
b = 10.7566 (18) Åθ = 2.8–22.4°
c = 9.964 (2) ŵ = 19.29 mm1
β = 100.557 (15)°T = 293 K
V = 415.47 (13) Å3Block, colourless
Z = 20.18 × 0.07 × 0.01 mm
Data collection top
Xcalibur, Ruby, Gemini R Ultra
diffractometer
323 reflections with I > 2σ(I)
ω–scanRint = 0.105
Absorption correction: gaussian
(CrysAlisPro; Rigaku OD, 2016) and (ABSORB; Angel et al., 2007)
θmax = 28.1°, θmin = 2.8°
Tmin = 0.361, Tmax = 0.434h = 55
2177 measured reflectionsk = 1211
550 independent reflectionsl = 1010
Refinement top
Refinement on F20 restraints
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.101H-atom parameters constrained
wR(F2) = 0.347 w = 1/[σ2(Fo2) + (0.2P)2]
where P = (Fo2 + 2Fc2)/3
S = 1.19(Δ/σ)max = 0.089
550 reflectionsΔρmax = 2.65 e Å3
46 parametersΔρmin = 2.89 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Br10.5857 (8)0.7915 (3)0.5946 (4)0.0384 (13)
Br20.3478 (7)0.5745 (3)0.8014 (4)0.0363 (13)
C10.540 (6)0.607 (5)0.555 (6)0.09 (2)
C20.430 (7)0.535 (3)0.630 (4)0.024 (8)
C30.404 (7)0.416 (3)0.587 (4)0.040 (10)
H30.3547500.3511950.6425770.048*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Br10.053 (2)0.017 (3)0.046 (4)0.0029 (13)0.0113 (19)0.0039 (12)
Br20.047 (2)0.023 (3)0.042 (4)0.0010 (14)0.0166 (19)0.0033 (12)
C10.004 (12)0.08 (4)0.18 (6)0.008 (17)0.01 (2)0.12 (4)
C20.023 (14)0.02 (3)0.04 (3)0.002 (13)0.019 (16)0.002 (12)
C30.022 (14)0.05 (3)0.04 (4)0.023 (15)0.003 (16)0.011 (16)
Geometric parameters (Å, º) top
Br1—C12.03 (4)C1—C3i1.49 (6)
Br2—C21.84 (3)C2—C31.36 (4)
C1—C21.20 (7)C3—H30.9300
C2—C1—C3i128 (3)C3—C2—Br2120 (2)
C2—C1—Br1122 (3)C2—C3—C1i116 (3)
C3i—C1—Br1109 (4)C2—C3—H3122.0
C1—C2—C3115 (4)C1i—C3—H3122.0
C1—C2—Br2125 (3)
C3i—C1—C2—C311 (6)Br1—C1—C2—Br210 (5)
Br1—C1—C2—C3178 (2)C1—C2—C3—C1i10 (5)
C3i—C1—C2—Br2176 (3)Br2—C2—C3—C1i177 (2)
Symmetry code: (i) x+1, y+1, z+1.
1,2,4,5-tetrabromobenzene (Mo-Absorb7_e) top
Crystal data top
C6H2Br4F(000) = 356
Mr = 393.72Dx = 3.147 Mg m3
Monoclinic, P21/nMo Kα radiation, λ = 0.71073 Å
a = 3.9431 (5) ÅCell parameters from 514 reflections
b = 10.7566 (18) Åθ = 2.8–22.4°
c = 9.964 (2) ŵ = 19.29 mm1
β = 100.557 (15)°T = 293 K
V = 415.47 (13) Å3Block, colourless
Z = 20.18 × 0.07 × 0.01 mm
Data collection top
Xcalibur, Ruby, Gemini R Ultra
diffractometer
313 reflections with I > 2σ(I)
ω–scanRint = 0.103
Absorption correction: gaussian
(CrysAlisPro; Rigaku OD, 2016) and (ABSORB; Angel et al., 2007)
θmax = 28.1°, θmin = 2.8°
Tmin = 0.361, Tmax = 0.434h = 55
2116 measured reflectionsk = 1211
531 independent reflectionsl = 1010
Refinement top
Refinement on F20 restraints
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.071H-atom parameters constrained
wR(F2) = 0.169 w = 1/[σ2(Fo2) + (0.0743P)2]
where P = (Fo2 + 2Fc2)/3
S = 1.05(Δ/σ)max = 0.592
531 reflectionsΔρmax = 1.04 e Å3
46 parametersΔρmin = 0.89 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Br10.5856 (4)0.79126 (19)0.5942 (2)0.0406 (8)
Br20.3480 (4)0.57453 (18)0.8014 (2)0.0377 (7)
C10.537 (3)0.6206 (17)0.541 (2)0.026 (5)
C20.440 (4)0.5357 (17)0.632 (2)0.030 (5)
C30.405 (4)0.4135 (14)0.5861 (18)0.020 (4)
H30.3410020.3529300.6431770.024*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Br10.0565 (13)0.021 (2)0.046 (2)0.0034 (9)0.0131 (12)0.0039 (8)
Br20.0503 (12)0.032 (2)0.034 (2)0.0011 (9)0.0154 (11)0.0028 (7)
C10.022 (8)0.010 (19)0.04 (2)0.008 (7)0.001 (9)0.000 (7)
C20.031 (9)0.000 (19)0.07 (2)0.006 (8)0.023 (11)0.003 (7)
C30.035 (9)0.006 (15)0.020 (18)0.003 (7)0.009 (9)0.011 (6)
Geometric parameters (Å, º) top
Br1—C11.911 (19)C1—C21.39 (2)
Br2—C21.84 (2)C2—C31.39 (2)
C1—C3i1.38 (3)C3—H30.9300
C3i—C1—C2122.4 (19)C3—C2—Br2119.3 (13)
C3i—C1—Br1119.1 (12)C1i—C3—C2121.9 (14)
C2—C1—Br1118.5 (17)C1i—C3—H3119.0
C1—C2—C3116 (2)C2—C3—H3119.0
C1—C2—Br2125.0 (17)
C3i—C1—C2—C30 (3)Br1—C1—C2—Br20.5 (18)
Br1—C1—C2—C3179.2 (10)C1—C2—C3—C1i0 (2)
C3i—C1—C2—Br2178.5 (13)Br2—C2—C3—C1i178.6 (13)
Symmetry code: (i) x+1, y+1, z+1.
1,2,4,5-tetrabromobenzene (Mo-CAsphere_f) top
Crystal data top
C6H2Br4F(000) = 356
Mr = 393.72Dx = 3.147 Mg m3
Monoclinic, P21/nMo Kα radiation, λ = 0.71073 Å
a = 3.9431 (5) ÅCell parameters from 514 reflections
b = 10.7566 (18) Åθ = 2.8–22.4°
c = 9.964 (2) ŵ = 19.29 mm1
β = 100.557 (15)°T = 293 K
V = 415.47 (13) Å3Block, colourless
Z = 20.18 × 0.07 × 0.01 × 0.03 (radius) mm
Data collection top
Xcalibur, Ruby, Gemini R Ultra
diffractometer
319 reflections with I > 2σ(I)
ω–scanRint = 0.102
Absorption correction: for a sphere
(CrysAlisPro; Rigaku OD, 2016)
θmax = 28.1°, θmin = 2.8°
Tmin = 0.638, Tmax = 0.645h = 55
2125 measured reflectionsk = 1211
531 independent reflectionsl = 1010
Refinement top
Refinement on F20 restraints
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.069H-atom parameters constrained
wR(F2) = 0.157 w = 1/[σ2(Fo2) + (0.0698P)2]
where P = (Fo2 + 2Fc2)/3
S = 1.05(Δ/σ)max = 0.523
531 reflectionsΔρmax = 0.93 e Å3
46 parametersΔρmin = 0.83 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Br10.5853 (4)0.79122 (17)0.5942 (2)0.0412 (7)
Br20.3482 (4)0.57458 (17)0.8016 (2)0.0387 (7)
C10.537 (3)0.6200 (15)0.5426 (19)0.024 (4)
C20.436 (4)0.5349 (16)0.630 (2)0.030 (5)
C30.406 (3)0.4125 (14)0.5870 (17)0.023 (4)
H30.3466320.3515910.6448340.027*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Br10.0580 (12)0.0185 (19)0.049 (2)0.0036 (8)0.0148 (11)0.0039 (7)
Br20.0505 (11)0.0301 (18)0.039 (2)0.0013 (8)0.0170 (10)0.0032 (7)
C10.020 (7)0.011 (17)0.04 (2)0.006 (7)0.003 (8)0.007 (7)
C20.033 (8)0.000 (17)0.06 (2)0.004 (7)0.022 (10)0.000 (7)
C30.031 (8)0.010 (15)0.029 (17)0.000 (7)0.008 (8)0.017 (6)
Geometric parameters (Å, º) top
Br1—C11.912 (17)C1—C3i1.40 (2)
Br2—C21.860 (18)C2—C31.38 (2)
C1—C21.37 (2)C3—H30.9300
C2—C1—C3i122.2 (17)C3—C2—Br2118.7 (12)
C2—C1—Br1120.1 (16)C2—C3—C1i120.2 (13)
C3i—C1—Br1117.7 (12)C2—C3—H3119.9
C1—C2—C3117.6 (18)C1i—C3—H3119.9
C1—C2—Br2123.7 (15)
C3i—C1—C2—C33 (2)Br1—C1—C2—Br21.2 (18)
Br1—C1—C2—C3179.9 (10)C1—C2—C3—C1i2 (2)
C3i—C1—C2—Br2178.7 (11)Br2—C2—C3—C1i178.7 (11)
Symmetry code: (i) x+1, y+1, z+1.
1,2,4,5-tetrabromobenzene (Mo-Absorb7-raw-Ciso_g) top
Crystal data top
C6H2Br4F(000) = 356
Mr = 393.72Dx = 3.147 Mg m3
Monoclinic, P21/nMo Kα radiation, λ = 0.71073 Å
a = 3.9431 (5) ÅCell parameters from 514 reflections
b = 10.7566 (18) Åθ = 2.8–22.4°
c = 9.964 (2) ŵ = 19.29 mm1
β = 100.557 (15)°T = 293 K
V = 415.47 (13) Å3Block, colourless
Z = 20.18 × 0.07 × 0.01 mm
Data collection top
Xcalibur, Ruby, Gemini R Ultra
diffractometer
323 reflections with I > 2σ(I)
ω–scanRint = 0.105
Absorption correction: gaussian
(CrysAlisPro; Rigaku OD, 2016) and (ABSORB; Angel et al., 2007)
θmax = 28.1°, θmin = 2.8°
Tmin = 0.361, Tmax = 0.434h = 55
2177 measured reflectionsk = 1211
550 independent reflectionsl = 1010
Refinement top
Refinement on F20 restraints
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.097H-atom parameters constrained
wR(F2) = 0.345 w = 1/[σ2(Fo2) + (0.2P)2]
where P = (Fo2 + 2Fc2)/3
S = 1.17(Δ/σ)max < 0.001
550 reflectionsΔρmax = 2.65 e Å3
31 parametersΔρmin = 2.90 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Br10.5856 (7)0.7911 (3)0.5945 (4)0.0376 (13)
Br20.3477 (7)0.5745 (3)0.8014 (4)0.0368 (13)
C10.539 (7)0.617 (3)0.545 (4)0.034 (7)*
C20.440 (6)0.535 (3)0.632 (3)0.022 (6)*
C30.404 (7)0.412 (3)0.587 (4)0.034 (7)*
H30.3438430.3505040.6435800.041*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Br10.053 (2)0.016 (3)0.044 (3)0.0026 (13)0.0115 (18)0.0036 (12)
Br20.046 (2)0.025 (3)0.042 (4)0.0015 (13)0.0158 (18)0.0030 (11)
Geometric parameters (Å, º) top
Br1—C11.94 (3)C1—C3i1.41 (5)
Br2—C21.84 (3)C2—C31.39 (4)
C1—C21.34 (4)C3—H30.9300
C2—C1—C3i125 (3)C3—C2—Br2119 (2)
C2—C1—Br1120 (3)C2—C3—C1i119 (3)
C3i—C1—Br1115 (2)C2—C3—H3120.5
C1—C2—C3116 (3)C1i—C3—H3120.5
C1—C2—Br2125 (3)
C3i—C1—C2—C32 (5)Br1—C1—C2—Br21 (3)
Br1—C1—C2—C3178.5 (18)C1—C2—C3—C1i2 (5)
C3i—C1—C2—Br2177 (2)Br2—C2—C3—C1i177 (2)
Symmetry code: (i) x+1, y+1, z+1.
1,2,4,5-tetrabromobenzene (Mo-Absorb7-Ciso_h) top
Crystal data top
C6H2Br4F(000) = 356
Mr = 393.72Dx = 3.147 Mg m3
Monoclinic, P21/nMo Kα radiation, λ = 0.71073 Å
a = 3.9431 (5) ÅCell parameters from 514 reflections
b = 10.7566 (18) Åθ = 2.8–22.4°
c = 9.964 (2) ŵ = 19.29 mm1
β = 100.557 (15)°T = 293 K
V = 415.47 (13) Å3Block, colourless
Z = 20.18 × 0.07 × 0.01 mm
Data collection top
Xcalibur, Ruby, Gemini R Ultra
diffractometer
313 reflections with I > 2σ(I)
ω–scanRint = 0.103
Absorption correction: gaussian
(CrysAlisPro; Rigaku OD, 2016) and (ABSORB; Angel et al., 2007)
θmax = 28.1°, θmin = 2.8°
Tmin = 0.361, Tmax = 0.434h = 55
2116 measured reflectionsk = 1211
531 independent reflectionsl = 1010
Refinement top
Refinement on F20 restraints
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.073H-atom parameters constrained
wR(F2) = 0.177 w = 1/[σ2(Fo2) + (0.0807P)2]
where P = (Fo2 + 2Fc2)/3
S = 1.04(Δ/σ)max < 0.001
531 reflectionsΔρmax = 1.03 e Å3
31 parametersΔρmin = 0.88 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Br10.5856 (4)0.79134 (18)0.5943 (2)0.0404 (8)
Br20.3480 (4)0.57458 (18)0.8015 (2)0.0376 (8)
C10.536 (3)0.6214 (16)0.5414 (19)0.026 (4)*
C20.440 (4)0.5358 (17)0.6328 (19)0.026 (4)*
C30.405 (3)0.4134 (15)0.5857 (19)0.022 (4)*
H30.3381500.3527640.6422730.026*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Br10.0565 (13)0.021 (2)0.045 (2)0.0031 (9)0.0129 (12)0.0038 (8)
Br20.0502 (12)0.031 (2)0.034 (2)0.0014 (9)0.0154 (11)0.0029 (7)
Geometric parameters (Å, º) top
Br1—C11.902 (18)C1—C21.40 (2)
Br2—C21.831 (18)C2—C31.40 (2)
C1—C3i1.38 (2)C3—H30.9300
C3i—C1—C2122.1 (18)C1—C2—Br2124.9 (15)
C3i—C1—Br1119.2 (12)C1i—C3—C2122.7 (15)
C2—C1—Br1118.7 (15)C1i—C3—H3118.6
C3—C2—C1115.2 (18)C2—C3—H3118.6
C3—C2—Br2119.9 (12)
C3i—C1—C2—C31 (2)Br1—C1—C2—Br21.3 (18)
Br1—C1—C2—C3179.1 (10)C1—C2—C3—C1i1 (2)
C3i—C1—C2—Br2178.6 (12)Br2—C2—C3—C1i178.8 (12)
Symmetry code: (i) x+1, y+1, z+1.
1,2,4,5-tetrabromobenzene (Mo-CAsphere-Ciso_i) top
Crystal data top
C6H2Br4F(000) = 356
Mr = 393.72Dx = 3.147 Mg m3
Monoclinic, P21/nMo Kα radiation, λ = 0.71073 Å
a = 3.9431 (5) ÅCell parameters from 514 reflections
b = 10.7566 (18) Åθ = 2.8–22.4°
c = 9.964 (2) ŵ = 19.29 mm1
β = 100.557 (15)°T = 293 K
V = 415.47 (13) Å3Block, colourless
Z = 20.18 × 0.07 × 0.01 × 0.03 (radius) mm
Data collection top
Xcalibur, Ruby, Gemini R Ultra
diffractometer
319 reflections with I > 2σ(I)
ω–scanRint = 0.102
Absorption correction: for a sphere
(CrysAlisPro; Rigaku OD, 2016)
θmax = 28.1°, θmin = 2.8°
Tmin = 0.638, Tmax = 0.645h = 55
2125 measured reflectionsk = 1211
531 independent reflectionsl = 1010
Refinement top
Refinement on F20 restraints
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.071H-atom parameters constrained
wR(F2) = 0.167 w = 1/[σ2(Fo2) + (0.078P)2]
where P = (Fo2 + 2Fc2)/3
S = 1.03(Δ/σ)max < 0.001
531 reflectionsΔρmax = 0.95 e Å3
31 parametersΔρmin = 0.82 e Å3
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Br10.5854 (4)0.79127 (17)0.5942 (2)0.0411 (7)
Br20.3482 (4)0.57464 (17)0.8016 (2)0.0388 (7)
C10.536 (3)0.6208 (15)0.5421 (18)0.024 (4)*
C20.438 (3)0.5350 (15)0.6310 (17)0.026 (4)*
C30.405 (3)0.4119 (14)0.5867 (18)0.024 (4)*
H30.3423700.3510160.6438100.029*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Br10.0581 (12)0.0182 (19)0.049 (2)0.0033 (8)0.0145 (11)0.0038 (7)
Br20.0505 (11)0.0300 (18)0.039 (2)0.0015 (8)0.0168 (10)0.0033 (7)
Geometric parameters (Å, º) top
Br1—C11.906 (17)C1—C3i1.39 (2)
Br2—C21.848 (17)C2—C31.39 (2)
C1—C21.38 (2)C3—H30.9300
C2—C1—C3i122.4 (17)C3—C2—Br2119.2 (11)
C2—C1—Br1119.6 (14)C1i—C3—C2120.7 (14)
C3i—C1—Br1117.9 (11)C1i—C3—H3119.6
C1—C2—C3116.9 (16)C2—C3—H3119.6
C1—C2—Br2123.9 (14)
C3i—C1—C2—C31 (2)Br1—C1—C2—Br20.0 (17)
Br1—C1—C2—C3179.7 (9)C1—C2—C3—C1i1 (2)
C3i—C1—C2—Br2178.8 (11)Br2—C2—C3—C1i178.8 (11)
Symmetry code: (i) x+1, y+1, z+1.
Comparison of technical characteristics of the diffractometers top
XtaLAB Synergy-S DualflexOxford Diffraction Gemini R Ultra
Radiation typeAg KαMo Kα
X-ray source typePhotonJet-S sourceEnhance X-ray source
Beam characteristics0.12 mm beam0.5 mm beam
X-ray opticsdouble-bounce multilayer opticsgraphite monochromator
Detector modelPilatus3 X CdTe 300KRuby
Detector typeHPC – photon countingCCD – integrative detector
Quantum efficiency>90%>80%
Read-out frequency (Hz)20<0.3
Goniometer4-circle Kappa goniometer (new generation)4-circle Kappa goniometer
Data-collection modeshutterless data collectionshuttered data collection
 

Acknowledgements

We are grateful to Dr Oleg Korneychik and TechnoInfo Ltd (Moscow, Russian Federation) for help with arranging the test diffraction experiment using the XtaLAB Synergy-S Dualflex diffractometer. We also thank Dr Ma­thias Meyer for technical help and discussions and Mr Adam Michalchuk for language polishing

Funding information

BAZ is grateful to the Russian Foundation for Basic Research (RFBR) for the financial support of research project No. 16–33-60093 mol_a_dk.

References

First citationAhsbahs, H. (2004). Z. Kristallogr. 219, 305–308.  Web of Science CrossRef CAS Google Scholar
First citationAngel, R. J., Bujak, M., Zhao, J., Gatta, G. D. & Jacobsen, S. D. (2007). J. Appl. Cryst. 40, 26–32.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationAngel, R. & Gonzalez-Platas, J. (2013). J. Appl. Cryst. 46, 252–254.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationBoehler, R. (2006). Rev. Sci. Instrum. 77, 115103; doi: 10.1063/1.2372734.  CrossRef Google Scholar
First citationBoldyreva, E. V. (2008). Acta Cryst. A64, 218–231.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationBoldyreva, E. V., Zakharov, B. A., Rashchenko, S. V., Seryotkin, Y. V. & Tumanov, N.A. (2016). Studying Solid-State Transformations Using In Situ X-Ray Diffraction Studies at High-Pressures. Novosibirsk: Publishing House of Siberian Branch of Russian Academy of Sciences, ISBN 978-5-7692-1526-1.  Google Scholar
First citationCasati, N., Genoni, A., Meyer, B., Krawczuk, A. & Macchi, P. (2017). Acta Cryst. B73, 584–597.  Web of Science CrossRef IUCr Journals Google Scholar
First citationCasati, N., Kleppe, A., Jephcoat, A. P. & Macchi, P. (2016). Nat. Commun. 7, 10901.  CSD CrossRef Google Scholar
First citationCasati, N., Macchi, P. & Sironi, A. (2007). J. Appl. Cryst. 40, 628–630.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationDawson, A., Allan, D. R., Parsons, S. & Ruf, M. (2004). J. Appl. Cryst. 37, 410–416.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationDera, P., Zhuravlev, K., Prakapenka, V., Rivers, M. L., Finkelstein, G. J., Grubor-Urosevic, O., Tschauner, O., Clark, S. M. & Downs, R. T. (2013). High. Press. Res. 33, 466–484.  Web of Science CrossRef CAS Google Scholar
First citationDubrovinsky, L., Boffa-Ballaran, T., Glazyrin, K., Kurnosov, A., Frost, D., Merlini, M., Hanfland, M., Prakapenka, V. B., Schouwink, P., Pippinger, T. & Dubrovinskaia, N. (2010). High. Press. Res. 30, 620–633.  CrossRef CAS Google Scholar
First citationForman, R. A., Piermarini, G. J., Dean Barnett, J. & Block, S. (1972). Science, 176, 284–285.  CrossRef PubMed CAS Web of Science Google Scholar
First citationGranero-García, R., Falenty, A. & Fabbiani, F. P. A. (2017). Chem. Eur. J. 23, 3691–3698.  Google Scholar
First citationHübschle, C. B., Sheldrick, G. M. & Dittrich, B. (2011). J. Appl. Cryst. 44, 1281–1284.  Web of Science CrossRef IUCr Journals Google Scholar
First citationKantor, I., Prakapenka, V., Kantor, A., Dera, P., Kurnosov, A., Sinogeikin, S., Dubrovinskaia, N. & Dubrovinsky, L. (2012). Rev. Sci. Instrum. 83, 125102.  Web of Science CrossRef PubMed Google Scholar
First citationKatrusiak, A. (1991). Cryst. Res. Technol. 26, 523–531.  CrossRef CAS Web of Science Google Scholar
First citationKatrusiak, A. (2004). Z. Kristallogr. 219, 461–467.  Web of Science CrossRef CAS Google Scholar
First citationKatrusiak, A. (2008). Acta Cryst. A64, 135–148.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMacrae, C. F., Bruno, I. J., Chisholm, J. A., Edgington, P. R., McCabe, P., Pidcock, E., Rodriguez-Monge, L., Taylor, R., van de Streek, J. & Wood, P. A. (2008). J. Appl. Cryst. 41, 466–470.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationMoggach, S. A., Allan, D. R., Parsons, S. & Warren, J. E. (2008). J. Appl. Cryst. 41, 249–251.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationParois, P., Moggach, S. A., Sanchez-Benitez, J., Kamenev, K. V., Lennie, A. R., Warren, J. E., Brechin, E. K., Parsons, S. & Murrie, M. (2010). Chem. Commun. 46, 1881–1883.  Web of Science CSD CrossRef CAS Google Scholar
First citationPiermarini, G. J., Block, S. & Barnett, J. D. (1973). J. Appl. Phys. 44, 5377–5382.  CrossRef CAS Web of Science Google Scholar
First citationPiermarini, G. J., Block, S., Barnett, J. D. & Forman, R. A. (1975). J. Appl. Phys. 46, 2774–2780.  CrossRef CAS Web of Science Google Scholar
First citationResnati, G., Boldyreva, E., Bombicz, P. & Kawano, M. (2015). IUCrJ, 2, 675–690.  Web of Science CrossRef CAS PubMed IUCr Journals Google Scholar
First citationRigaku OD (2016). CrysAlis PRO. Rigaku Oxford Diffraction Ltd, Yarnton, England.  Google Scholar
First citationSahoo, S. C., Sinha, S. B., Kiran, M. S. R. N., Ramamurty, U., Dericioglu, A. F., Reddy, C. M. & Naumov, P. (2013). J. Am. Chem. Soc. 135, 13843–13850.  Web of Science CrossRef CAS PubMed Google Scholar
First citationSaouane, S. & Fabbiani, F. P. A. (2015). Cryst. Growth Des. 15, 3875–3884.  CSD CrossRef CAS Google Scholar
First citationSaouane, S., Norman, S. E., Hardacre, C. & Fabbiani, F. P. A. (2013). Chem. Sci. 4, 1270–1280.  Web of Science CSD CrossRef CAS Google Scholar
First citationSheldrick, G. M. (2015). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSowa, H. & Ahsbahs, H. (2006). J. Appl. Cryst. 39, 169–175.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSpek, A. L. (2009). Acta Cryst. D65, 148–155.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationVeciana, J., Souto, M., Gullo, M. C., Cui, H., Casati, N., Montisci, F., Jeschke, H. O., Valentí, R., Ratera, I. & Rovira, C. (2018). Chem. Eur. J., doi 10.1002cem.201800881.  Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationYan, H., Yang, F., Pan, D., Lin, Y., Hohman, J. N., Solis-Ibarra, D., Li, F. H., Dahl, J. E. P., Carlson, R. M. K., Tkachenko, B. A., Fokin, A. A., Schreiner, P. R., Galli, G., Mao, W. L., Shen, Z.-X. & Melosh, N. A. (2018). Nature, 554, 505–510.  CSD CrossRef CAS Google Scholar
First citationZakharov, B. A., Gribov, P. A., Matvienko, A. A. & Boldyreva, E. V. (2017). Z. Kristallogr. 232, 751–757.  CAS Google Scholar
First citationZakharov, B. A., Michalchuk, A. A. L., Morrison, C. A. & Boldyreva, E. V. (2018). Phys. Chem. Chem. Phys., doi: 10.1039/C7CP08609A.  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.

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890
Follow Acta Cryst. E
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds