research communications\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Crystal structure of 2,5-di­hy­droxy­terephthalic acid from powder diffraction data

crossmark logo

aDepartment of Chemistry, North Central College, 131 S. Loomis, St., Naperville IL, 60540 , USA
*Correspondence e-mail: kaduk@polycrystallography.com

Edited by W. T. A. Harrison, University of Aberdeen, Scotland (Received 16 September 2022; accepted 25 September 2022; online 30 September 2022)

The crystal structure of anhydrous 2,5-dhy­droxy­terephthalic acid, C8H6O6, was solved and refined using laboratory X-ray powder diffraction data, and optimized using density functional techniques. The published structure of 2,5-di­hydroxy­terephthalic acid dihydrate was also optimized. The carb­oxy­lic acid groups form strong hydrogen bonds, which form centrosymmetric rings with graph set R22(8). These hydrogen bonds link the mol­ecules into chains along [011]. There is an intra­molecular O—H⋯O hydrogen bond between the hydroxyl group and the carbonyl group of the carb­oxy­lic acid. The hydrogen bonding in the dihydrate is very different. Although the intra­molecular hy­droxy/carb­oxy­lic acid hydrogen bond is present, the water mol­ecule acts as an acceptor to the carb­oxy­lic acid and a donor to two other oxygen atoms. The carb­oxy­lic acid groups do not inter­act with each other directly.

1. Chemical context

2,5-Di­hydroxy­terephthalate (C8H4O62–; dhtp) is of current inter­est as a linker in metal–organic frameworks (MOFs). It can add extra functionality to alter adsorption and catalytic properties. In an attempt to replicate the ionothermal preparation of the Co-dhtp MOF Co2(dobdc)-ST (Azbell et al., 2022[Azbell, T., Pitt, T., Bollmeyer, M., Cong, C., Lancaster, K. & Milner, P. (2022). ChemRxiv, https://doi.org/10.26434/chemrxiv-2022-00xd7.]), an unexpected product was obtained, namely anhydrous 2,5-dhy­droxy­terephthalic acid, C8H6O6, (I)[link].

[Scheme 1]

The crystal structures of three Co-dhtp MOFs have been reported: Cambridge Structural Database refcodes FEGBEB (Gen, 2017[Gen, Z. L. (2017). Private communication (refcode FEGBEB). CCDC, Cambridge, England. https://doi.org/10.5517/ccdc.csd.cc1pfslh]), VOFJIM (Rosnes et al., 2019[Rosnes, M. H., Mathieson, J. S., Törnroos, K. W., Johnsen, R. E., Cronin, L. & Dietzel, P. D. C. (2019). Cryst. Growth Des. 19, 2089-2096.]) and VOFJIM01 (Ayoub et al., 2019[Ayoub, G., Karadeniz, B., Howarth, A. J., Farha, O. K., Đilović, I., Germann, L. S., Dinnebier, R. E., Užarević, K. & Friščić, T. (2019). Chem. Mater. 31, 5494-5501.]). The calculated powder patterns of these three compounds, which have been given the name CPO-27-Co, indicate that they have the same structure (Fig. 1[link]).

[Figure 1]
Figure 1
Calculated (using Mercury; Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]) powder diffraction patterns (Cu Kα radiation) for CPO-27-Co [FEGBEB (Gen, 2017[Gen, Z. L. (2017). Private communication (refcode FEGBEB). CCDC, Cambridge, England. https://doi.org/10.5517/ccdc.csd.cc1pfslh]), VOFJIM (Rosnes et al., 2019[Rosnes, M. H., Mathieson, J. S., Törnroos, K. W., Johnsen, R. E., Cronin, L. & Dietzel, P. D. C. (2019). Cryst. Growth Des. 19, 2089-2096.]) and VOFJIM01 (Ayoub et al., 2019[Ayoub, G., Karadeniz, B., Howarth, A. J., Farha, O. K., Đilović, I., Germann, L. S., Dinnebier, R. E., Užarević, K. & Friščić, T. (2019). Chem. Mater. 31, 5494-5501.])]. The differences in peak positions result from the different temperatures of the diffraction studies. Image generated using JADE Pro (MDI, 2022[MDI (2022). JADE Pro. Materials Data, Livermore, CA, USA.]).

2. Structural commentary

Compound (I)[link] crystallizes in the triclinic space group P[\overline{1}] with half a mol­ecule in the asymmetric unit. The root-mean-square Cartesian displacements of the non-H atoms in the Rietveld-refined and CRYSTAL17-optimized structures is 0.053 Å (Fig. 2[link]). The good agreement provides strong evidence that the structure is correct (van de Streek & Neumann, 2014[Streek, J. van de & Neumann, M. A. (2014). Acta Cryst. B70, 1020-1032.]). The CRYSTAL17 and VASP-optimized structures are essentially identical (r.m.s. displacement = 0.031 Å). This discussion concentrates on the CRYSTAL17-optimized structure. The full mol­ecule (with atom numbering) is illustrated in Fig. 3[link] and a view of the packing down the a-axis direction is shown in Fig. 4[link].

[Figure 2]
Figure 2
Comparison of the Rietveld-refined (red) and VASP-optimized (blue) structures of anhydrous 2,5-di­hydroxy­terephthalic acid. The r.m.s. Cartesian displacement is 0.053 Å. Image generated using Mercury (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]).
[Figure 3]
Figure 3
The full 2,5-di­hydroxy­terephthalic acid mol­ecule, with the atom numbering. The atoms are represented by 50% probability spheroids. Image generated using Mercury (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]). Symmetry code: (a) 1 − x, 1 − y, 1 − z.
[Figure 4]
Figure 4
The crystal structure of anhydrous 2,5-di­hydroxy­terephthalic acid, viewed down the a-axis. Image generated using DIAMOND (Crystal Impact, 2022[Crystal Impact (2022). DIAMOND. Crystal Impact GbR, Bonn, Germany. https://www.crystalimpact.de/diamond]).

All of the bond distances, angles, and torsion angles fall within the normal ranges indicated by a Mercury Mogul geometry check (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]). The plane of the phenyl ring lies approximately on the (98[\overline{9}]) Miller plane. The peak profiles are dominated by anisotropic microstrain broadening: the average microstrain is 8362 ppm.

The Bravais–Friedel–Donnay–Harker (Bravais, 1866[Bravais, A. (1866). Etudes Cristallographiques. Paris: Gauthier Villars.]; Friedel, 1907[Friedel, G. (1907). Bull. Soc. Fr. Mineral. 30, 326-455.]; Donnay & Harker, 1937[Donnay, J. D. H. & Harker, D. (1937). Am. Mineral. 22, 446-447.]) morphology suggests that we might expect platy (with {001} as the major faces) morphology for this crystal. A 4th order spherical harmonics preferred orientation model was included in the refinement. The refined texture index was 1.059 (2), indicating that preferred orientation was small for this capillary specimen. In flat plate specimens examined in Bragg–Brentano geometry using Cu radiation, the preferred orientation tended to be higher.

3. Supra­molecular features

In the extended structure of (I)[link], the carb­oxy­lic acid groups form strong O3—H4⋯O5 hydrogen bonds, which form centrosymmetric loops with graph set R22(8) (Etter, 1990[Etter, M. C. (1990). Acc. Chem. Res. 23, 120-126.]; Bernstein et al., 1995[Bernstein, J., Davis, R. E., Shimoni, L. & Chang, N. L. (1995). Angew. Chem. Int. Ed. Engl. 34, 1555-1573.]; Shields et al., 2000[Shields, G. P., Raithby, P. R., Allen, F. H. & Motherwell, W. D. S. (2000). Acta Cryst. B56, 455-465.]). These hydrogen bonds link the mol­ecules into chains propagating along [011] (Table 1[link]; Fig. 5[link]). There is an intra­molecular O1—H2⋯O5 hydrogen bond between the hydroxyl group and the carbonyl group of the carboxyl acid. A C—H⋯O hydrogen bond also contributes to the lattice energy. The Mercury aromatics analyser indicates one strong inter­action with a centroid–centroid distance of 4.26 Å, and a moderate one at 5.59 Å.

Table 1
Hydrogen-bond geometry (Å, °) for (I)[link]

D—H⋯A D—H H⋯A DA D—H⋯A
O3—H4⋯O5i 1.00 1.69 2.689 174
O1—H2⋯O5 0.99 1.68 2.567 147
Symmetry code: (i) [-x+1, -y+2, -z+2].
[Figure 5]
Figure 5
The hydrogen bonds in the structure of anhydrous 2,5-di­hydroxy­terephthalic acid. Image generated using Mercury (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]).

The hydrogen bonding in the dihydrate DUSJUX (Cheng et al., 2010[Cheng, P.-W., Cheng, C.-F., Chun-Ting, Y. & Lin, C.-H. (2010). Acta Cryst. E66, o1928.]) is very different (Table 2[link]; Fig. 6[link]). Although the intra­molecular hy­droxy–carb­oxy­lic acid O—H⋯O hydrogen bond is present, the water mol­ecule acts as an acceptor to the carb­oxy­lic acid and a donor to two other oxygen atoms. The carb­oxy­lic acid groups do not inter­act with each other directly.

Table 2
Hydrogen-bond geometry (Å, °) for DUSJUX[link]

D—H⋯A D—H H⋯A DA D—H⋯A
O2—H2⋯O4 1.07 1.43 2.500 178
O1—H1⋯O3i 1.01 1.64 2.562 149
O4—H4⋯O3ii 0.99 1.78 2.736 161
O4—H5⋯O1iii 0.99 1.82 2.794 169
Symmetry codes: (i) [-x, -y+1, -z+2]; (ii) [x, -y+{\script{3\over 2}}, z-{\script{1\over 2}}]; (iii) [-x+1, -y+1, -z+1].
[Figure 6]
Figure 6
The hydrogen bonds in the structure of 2–5-di­hydroxy­terephthalic acid dihydrate DUSJUX. Image generated using Mercury (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]).

The CRYSTAL17 (Dovesi et al., 2018[Dovesi, R., Erba, A., Orlando, R., Zicovich-Wilson, C. M., Civalleri, B., Maschio, L., Rérat, M., Casassa, S., Baima, J., Salustro, J. & Kirtman, B. (2018). WIREs Comput. Mol. Sci. 8, e1360.]) calculations suggest that DUSJUX is 28.5 kcal mol−1 lower in energy than the sum of anhdyrous 2,5-di­hydroxy­terephthalic acid and two water mol­ecules. The corresponding VASP (Kresse & Furthmüller, 1996[Kresse, G. & Furthmüller, J. (1996). Comput. Mater. Sci. 6, 15-50.]) calculations indicate that DUSJUX is 114.0 kcal mol−1 more stable. As chemists, we would like to attribute the `extra' energy to the formation of additional hydrogen bonds. Rammohan & Kaduk (2018[Rammohan, A. & Kaduk, J. A. (2018). Acta Cryst. B74, 239-252.]) developed (for citrates using earlier versions of CRYSTAL) a correlation between the energy of an O—H⋯O hydrogen bond and the Mulliken overlap population between the H and the O acceptor: E (kcal mol−1) = 54.7(overlap)1/2. Using this correlation to estimate the energies of the individual hydrogen bonds, we calculate that DUSJUX is 59.6 kcal mol−1 lower in energy than the sum of the anhydrous mol­ecule and two water mol­ecules – a value between the two DFT calculations. While the disagreements indicate that the absolute energy calculated for a hydrogen bond may be more uncertain than we would like, the Mulliken overlap population is certainly a guide to whether a hydrogen bond is stronger or weaker than another, and to whether a (geometrically possible) hydrogen bond is real or not.

4. Database survey

A connectivity search in the Cambridge Structural Database [CSD version 5.43 June 2022 (Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]); ConQuest 2022.2.0 (Bruno et al., 2002[Bruno, I. J., Cole, J. C., Edgington, P. R., Kessler, M., Macrae, C. F., McCabe, P., Pearson, J. & Taylor, R. (2002). Acta Cryst. B58, 389-397.])] of a 2,5-di­hydroxy­terephthalate fragment with the elements C, H, and O only yielded the structure of the dihydrate (Cheng et al., 2010[Cheng, P.-W., Cheng, C.-F., Chun-Ting, Y. & Lin, C.-H. (2010). Acta Cryst. E66, o1928.]; DUSJUX), as well as two esters. The dihydrate was also obtained accidentally during the synthesis of metal–organic coordination polymers. Removing the chemistry limitation yielded 249 entries, many of which are metal–organic frameworks. A search of the powder pattern against the Powder Diffraction File (Gates-Rector & Blanton, 2019[Gates-Rector, S. & Blanton, T. (2019). Powder Diffr. 34, 352-360.]) yielded no hits. Not even the usual accidental matches were obtained; this pattern evidently occupies a very different portion of `diffraction space'.

5. Synthesis and crystallization

Cobalt(II) chloride hexa­hydrate (1.78 g, 7.50 mmol) and 2,5-di­hydroxy­terephthalic acid (1.00 g, 5.05 mmol) were crushed together with mortar and pestle and added to a 10 ml round-bottom flask. The flask was connected to a Schlenk line and placed in a glass bowl of sand on top of a hot plate. The hot plate was heated to 443 K for approximately 18 h and the round-bottom flask was under vacuum. After being removed from the heat and allowed to cool, the remaining solid was transferred to a Pyrex container with aceto­nitrile (50 ml) and placed in a vacuum oven at 343 K for 24 h. After removal from the oven, the solution was deca­nted and replaced with aceto­nitrile (50 ml). This wash procedure was done a total of three times. The remaining solid was then added to 100 ml of methanol at 343 K for 24 h and deca­nted, this wash was done two times. The remaining solid was then added to a vacuum oven at 423 K for 24 h. The remaining solid was then added to a scintillation vial wrapped with Parafilm for storage.

6. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 3[link]. A portion of the sample was blended with 11.51% < 1 micron diamond powder (Alfa Aesar) inter­nal standard in a mortar and pestle until the color was uniform. The X-ray powder diffraction pattern was measured from a 0.7 mm diameter static capillary specimen on a PANalytical Empyrean diffractometer using Mo Kα radiation. The pattern was measured from 1.0–50.0° 2θ in 0.0083560° steps, counting for 4 sec/step.

Table 3
Experimental details

  (I) DUSJUX (DFT)
Crystal data
Chemical formula C8H6O6 C8H6O6·2(H2O)
Mr 198.08 --
Crystal system, space group Triclinic, P[\overline{1}] Monoclinic, P21/c
Temperature (K) 302 --
a, b, c (Å) 4.2947 (5), 5.6089 (5), 8.2331 (19) 5.18830, 17.54500, 5.49900
α, β, γ (°) 93.612 (4), 102.219 (4), 96.7621 (14) 90, 103.03, 90
V3) 191.69 (1) 487.68
Z 1 2
Radiation type Mo Kα1,2, λ = 0.70932, 0.71361 Å --
Specimen shape, size (mm) Cylinder, 12 × 0.7 --
 
Data collection
Diffractometer PANalytical Empyrean  
Specimen mounting Glass capillary  
Data collection mode Transmission  
Data collection method Step  
θ values (°) 2θmin = 1.002 2θmax = 49.991 2θstep = 0.008  
 
Refinement
R factors and goodness of fit Rp = 0.034, Rwp = 0.042, Rexp = 0.019, χ2 = 5.148  
No. of parameters 53  
No. of restraints 18  
(Δ/σ)max 2.635  
The same symmetry and lattice parameters were used for the DFT calculations as for the powder diffraction study for (I). Computer program: GSAS-II (Toby & Von Dreele, 2013[Rammohan, A. & Kaduk, J. A. (2018). Acta Cryst. B74, 239-252.]).

After correcting the peak positions using the known diamond peak positions, the pattern was indexed using JADE Pro (MDI, 2022[MDI (2022). JADE Pro. Materials Data, Livermore, CA, USA.]) on a primitive triclinic cell with a = 4.26420, b = 5.58601, c = 8.17902 Å, α = 93.53, β = 12.13, γ = 96.78° and V = 188 Å3. Since the volume corresponded to one mol­ecule of 2,5-di­hydroxy­terephthalic acid, the space group was assumed to be P[\overline{1}], with half a mol­ecule in the asymmetric unit. A reduced cell search of the CSD yielded no hits. Preliminary indexing attempts using the default peak list from a pattern collected using Cu radiation were unsuccessful (monoclinic cells with no reasonable structures), until closer examination of the pattern revealed that the peak at 21.6° (9.7° Mo) was actually a doublet, and that there was an additional peak at 22.0° (9.9° Mo). Including these two additional peaks yielded the triclinic cell.

The 2,5-di­hydroxy­terephthalic acid mol­ecule was extracted from the DUSJUX structure using Materials Studio (Dassault Systèmes, 2021[Dassault Systèmes (2021). Materials Studio. BIOVIA, San Diego, USA.]), and saved as a .mol2 file. The crystal structure was solved using Monte Carlo simulated annealing techniques as implemented in EXPO2014 (Altomare et al., 2013[Altomare, A., Cuocci, C., Giacovazzo, C., Moliterni, A., Rizzi, R., Corriero, N. & Falcicchio, A. (2013). J. Appl. Cryst. 46, 1231-1235.]), using a whole mol­ecule as the fragment. Since the mol­ecule occupies a center of symmetry, the two halves overlapped partially. The overlapping atoms were averaged manually using Materials Studio to obtain the asymmetric unit.

Rietveld refinement was carried out using GSAS-II (Toby & Von Dreele, 2013[Toby, B. H. & Von Dreele, R. B. (2013). J. Appl. Cryst. 46, 544-549.]). All non-H bond distances and angles were subjected to restraints, based on a Mercury Mogul geometry check (Sykes et al., 2011[Sykes, R. A., McCabe, P., Allen, F. H., Battle, G. M., Bruno, I. J. & Wood, P. A. (2011). J. Appl. Cryst. 44, 882-886.]; Bruno et al., 2004[Bruno, I. J., Cole, J. C., Kessler, M., Luo, J., Motherwell, W. D. S., Purkis, L. H., Smith, B. R., Taylor, R., Cooper, R. I., Harris, S. E. & Orpen, A. G. (2004). J. Chem. Inf. Comput. Sci. 44, 2133-2144.]). A planar restraint was applied to the benzene ring. The Mogul average and standard deviation for each qu­antity were used as the restraint parameters. The restraints contributed 1.9% to the final χ2. The hydrogen atoms were included in calculated positions, which were recalculated during the refinement using Materials Studio (Dassault Systèmes, 2021[Dassault Systèmes (2021). Materials Studio. BIOVIA, San Diego, USA.]). The Uiso of the heavy atoms were grouped by chemical similarity. The Uiso for the H atoms were fixed at 1.3× the Uiso of the heavy atoms to which they are attached. The peak profiles were described using the generalized microstrain model. The background was modeled using a four-term shifted Chebyshev polynomial, along with a peak at 12.05° to model the scattering from the glass capillary and any amorphous component. The final refinements yielded the residuals reported in Table 1[link]. The largest errors in the difference plot (Fig. 7[link]) are small, and are in the shapes of the peaks.

[Figure 7]
Figure 7
The Rietveld plot for the refinement of anhydrous 2,5-di­hydroxy­terephthalic acid. The blue crosses represent the observed data points, and the green line is the calculated pattern. The cyan curve is the normalized error plot, and the red line is the background curve. The row of tick marks indicates the calculated reflection positions. The vertical scale has been multiplied by a factor of 10× for 2θ > 20.5°. The row of red tick marks indicate the positions of the diamond internal standard peaks.

The crystal structure (as well as that of DUSJUX and an isolated water mol­ecule) was optimized using VASP (Kresse & Furthmüller, 1996[Kresse, G. & Furthmüller, J. (1996). Comput. Mater. Sci. 6, 15-50.]) (fixed experimental unit cells) through the MedeA graphical inter­face (Materials Design, 2016[Materials Design (2016). MedeA. Materials Design Inc., Angel Fire, NM, USA.]). The calculations were carried out on 16 2.4 GHz processors (each with 4 Gb RAM) of a 64-processor HP Proliant DL580 Generation 7 Linux cluster at North Central College. The calculations used the GGA-PBE functional, a plane wave cutoff energy of 400.0 eV, and a k-point spacing of 0.5 Å−1 leading to a 4 × 3 × 2 mesh. The structures were also optimized (fixed experimental cells) and population analyses were carried out using CRYSTAL17 (Dovesi et al., 2018[Dovesi, R., Erba, A., Orlando, R., Zicovich-Wilson, C. M., Civalleri, B., Maschio, L., Rérat, M., Casassa, S., Baima, J., Salustro, J. & Kirtman, B. (2018). WIREs Comput. Mol. Sci. 8, e1360.]). The basis sets for the H, C, N, and O atoms in the calculations were those of Gatti et al. (1994[Gatti, C., Saunders, V. R. & Roetti, C. (1994). J. Chem. Phys. 101, 10686-10696.]). The calculations were run on a 3.5 GHz PC using 8 k-points and the B3LYP functional.

Supporting information


Computing details top

2,5-Dihydroxybenzene-1,4-dicarboxylic acid (I) top
Crystal data top
C8H6O6β = 102.219 (4)°
Mr = 198.08γ = 96.7621 (14)°
Triclinic, P1V = 191.69 (1) Å3
Hall symbol: -P 1Z = 1
a = 4.2947 (5) ÅDx = 1.716 Mg m3
b = 5.6089 (5) ÅT = 302 K
c = 8.2331 (19) Åcylinder, 12 × 0.7 mm
α = 93.612 (4)°
Data collection top
PANalytical Empyrean
diffractometer
Data collection mode: transmission
Specimen mounting: glass capillaryScan method: step
Refinement top
18 restraintsPreferred orientation correction: Simple spherical harmonic correction Order = 4 Coefficients: 0:0:C(2,-2) = 0.246(11); 0:0:C(2,-1) = -0.018(11); 0:0:C(2,0) = -0.313(16); 0:0:C(2,1) = 0.217(13); 0:0:C(2,2) = -0.192(9); 0:0:C(4,-4) = -0.146(17); 0:0:C(4,-3) = 0.073(19); 0:0:C(4,-2) = -0.052(16); 0:0:C(4,-1) = 0.083(18); 0:0:C(4,0) = -0.058(17); 0:0:C(4,1) = -0.006(18); 0:0:C(4,2) = -0.196(23); 0:0:C(4,3) = 0.071(16); 0:0:C(4,4) = 0.108(25)
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
C100.492 (2)0.6406 (16)0.6392 (11)0.0323 (10)*
C60.6818 (18)0.4506 (18)0.6571 (9)0.0323 (10)*
C70.6946 (18)0.3106 (14)0.5107 (12)0.0323 (10)*
C90.477 (2)0.7892 (16)0.7918 (9)0.0553 (15)*
O10.8599 (12)0.4082 (9)0.8120 (6)0.0323 (10)*
O30.2791 (12)0.9640 (12)0.7718 (7)0.0553 (15)*
O50.6452 (16)0.7744 (11)0.9376 (9)0.0553 (15)*
H80.841850.171350.529280.0420 (14)*
H20.837820.542880.889980.0420 (14)*
H40.302801.072550.876630.0719 (19)*
Geometric parameters (Å, º) top
C10—C61.411 (5)C9—O31.365 (5)
C10—C7i1.384 (5)C9—O51.277 (5)
C10—C91.482 (6)O1—C61.391 (5)
C6—C101.411 (5)O1—H20.987 (5)
C6—C71.412 (6)O3—C91.365 (5)
C6—O11.391 (5)O3—H41.004 (6)
C7—C10i1.384 (5)O5—C91.277 (5)
C7—C61.412 (6)H8—C71.060 (8)
C7—H81.060 (8)H2—O10.987 (5)
C9—C101.482 (6)H4—O31.004 (6)
C6—C10—C7i124.3 (7)C10i—C7—H8126.7 (10)
C6—C10—C9118.1 (9)C6—C7—H8115.1 (10)
C7i—C10—C9117.5 (9)C10—C9—O3116.8 (7)
C10—C6—C7117.4 (6)C10—C9—O5125.0 (9)
C10—C6—O1121.6 (9)O3—C9—O5118.1 (7)
C7—C6—O1121.0 (10)C6—O1—H2105.5 (6)
C10i—C7—C6118.1 (7)C9—O3—H4112.7 (5)
Symmetry code: (i) x+1, y+1, z+1.
(Ia) top
Crystal data top
CV = 46.12 (1) Å3
Mr = 12.01Z = 8
Cubic, Fd3mDx = 3.459 Mg m3
Hall symbol: -F 4vw 2vwT = 302 K
a = 3.58625 (11) Å
Refinement top
Preferred orientation correction: March-Dollase correction coef. = 1.000 axis = [0, 0, 1]
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
C10.125000.125000.125000.0159*
Geometric parameters (Å, º) top
C1—C1i1.5529C1—C1iii1.5529
C1—C1ii1.5529C1—C1iv1.5529
C1i—C1—C1ii109.471C1i—C1—C1iv109.471
C1i—C1—C1iii109.471C1ii—C1—C1iv109.471
C1ii—C1—C1iii109.471C1iii—C1—C1iv109.471
Symmetry codes: (i) x+1/4, y+1/4, z; (ii) z, x+1/4, y+1/4; (iii) y+1/4, z, x+1/4; (iv) x, y, z.
(I_DFT) top
Crystal data top
C8H6O6c = 8.1976 Å
Mr = 198.08α = 93.6590°
Triclinic, P1β = 102.1730°
a = 4.2647 Åγ = 96.7840°
b = 5.5912 ÅZ = 1
Data collection top
h = l =
k =
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
C100.507250.651650.640680.06414*
C60.686950.463500.655710.06414*
C70.690970.313930.515460.06414*
C90.484030.806910.790690.01062*
O10.876560.424230.804580.01062*
O30.289420.974620.768300.01062*
O50.650170.780630.931090.01062*
H80.841850.171350.529280.08339*
H20.837820.542880.889990.01381*
H40.302801.072550.876630.01381*
Bond lengths (Å) top
C10—C61.370C7—H81.079
C10—C7i1.415C9—O31.320
C10—C91.487C9—O51.245
C6—C71.382O1—H20.986
C6—O11.361O3—H41.000
C7—C10i1.415H4—O31.000
Symmetry code: (i) x+1, y+1, z+1.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O3—H4···O5ii1.001.692.689174
O1—H2···O50.991.682.567147
Symmetry code: (ii) x+1, y+2, z+2.
(DUSJUX_DFT) top
Crystal data top
C8H6O6·2(H2O)c = 5.49900 Å
Monoclinic, P21/cβ = 103.03°
a = 5.18830 ÅV = 487.68 Å3
b = 17.54500 ÅZ = 2
Data collection top
h = l =
k =
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzBiso*/Beq
O10.028290.351000.84164
H10.082290.319150.93268
O20.418420.594300.63866
H20.512660.640440.56603
O30.281230.684920.87357
C10.011450.424010.92007
C20.146920.481060.82449
H30.262530.467070.68792
C30.285850.616630.80175
C40.138410.556820.90233
O40.646580.700110.46632
H40.525370.742630.39890
H50.751450.686800.34282
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O2—H2···O41.071.432.500178
O1—H1···O3i1.011.642.562149
O4—H4···O3ii0.991.782.736161
O4—H5···O1iii0.991.822.794169
Symmetry codes: (i) x, y+1, z+2; (ii) x, y+3/2, z1/2; (iii) x+1, y+1, z+1.
 

Acknowledgements

We thank Professors Nicholas C. Boaz, Paul F. Brandt and Jeffrey A. Bjorklund for guidance and helpful discusssions.

References

First citationAltomare, A., Cuocci, C., Giacovazzo, C., Moliterni, A., Rizzi, R., Corriero, N. & Falcicchio, A. (2013). J. Appl. Cryst. 46, 1231–1235.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationAyoub, G., Karadeniz, B., Howarth, A. J., Farha, O. K., Đilović, I., Germann, L. S., Dinnebier, R. E., Užarević, K. & Friščić, T. (2019). Chem. Mater. 31, 5494–5501.  CrossRef CAS Google Scholar
First citationAzbell, T., Pitt, T., Bollmeyer, M., Cong, C., Lancaster, K. & Milner, P. (2022). ChemRxiv, https://doi.org/10.26434/chemrxiv-2022-00xd7.  Google Scholar
First citationBernstein, J., Davis, R. E., Shimoni, L. & Chang, N. L. (1995). Angew. Chem. Int. Ed. Engl. 34, 1555–1573.  CrossRef CAS Web of Science Google Scholar
First citationBravais, A. (1866). Etudes Cristallographiques. Paris: Gauthier Villars.  Google Scholar
First citationBruno, I. J., Cole, J. C., Edgington, P. R., Kessler, M., Macrae, C. F., McCabe, P., Pearson, J. & Taylor, R. (2002). Acta Cryst. B58, 389–397.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationBruno, I. J., Cole, J. C., Kessler, M., Luo, J., Motherwell, W. D. S., Purkis, L. H., Smith, B. R., Taylor, R., Cooper, R. I., Harris, S. E. & Orpen, A. G. (2004). J. Chem. Inf. Comput. Sci. 44, 2133–2144.  Web of Science CrossRef PubMed CAS Google Scholar
First citationCheng, P.-W., Cheng, C.-F., Chun-Ting, Y. & Lin, C.-H. (2010). Acta Cryst. E66, o1928.  CrossRef IUCr Journals Google Scholar
First citationCrystal Impact (2022). DIAMOND. Crystal Impact GbR, Bonn, Germany. https://www.crystalimpact.de/diamond  Google Scholar
First citationDassault Systèmes (2021). Materials Studio. BIOVIA, San Diego, USA.  Google Scholar
First citationDonnay, J. D. H. & Harker, D. (1937). Am. Mineral. 22, 446–447.  CAS Google Scholar
First citationDovesi, R., Erba, A., Orlando, R., Zicovich–Wilson, C. M., Civalleri, B., Maschio, L., Rérat, M., Casassa, S., Baima, J., Salustro, J. & Kirtman, B. (2018). WIREs Comput. Mol. Sci. 8, e1360.  Google Scholar
First citationEtter, M. C. (1990). Acc. Chem. Res. 23, 120–126.  CrossRef CAS Web of Science Google Scholar
First citationFriedel, G. (1907). Bull. Soc. Fr. Mineral. 30, 326–455.  Google Scholar
First citationGates-Rector, S. & Blanton, T. (2019). Powder Diffr. 34, 352–360.  CAS Google Scholar
First citationGatti, C., Saunders, V. R. & Roetti, C. (1994). J. Chem. Phys. 101, 10686–10696.  CrossRef CAS Web of Science Google Scholar
First citationGen, Z. L. (2017). Private communication (refcode FEGBEB). CCDC, Cambridge, England. https://doi.org/10.5517/ccdc.csd.cc1pfslh  Google Scholar
First citationGroom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171–179.  Web of Science CrossRef IUCr Journals Google Scholar
First citationKresse, G. & Furthmüller, J. (1996). Comput. Mater. Sci. 6, 15–50.  CrossRef CAS Web of Science Google Scholar
First citationMacrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226–235.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMaterials Design (2016). MedeA. Materials Design Inc., Angel Fire, NM, USA.  Google Scholar
First citationMDI (2022). JADE Pro. Materials Data, Livermore, CA, USA.  Google Scholar
First citationRammohan, A. & Kaduk, J. A. (2018). Acta Cryst. B74, 239–252.  Web of Science CSD CrossRef IUCr Journals Google Scholar
First citationRosnes, M. H., Mathieson, J. S., Törnroos, K. W., Johnsen, R. E., Cronin, L. & Dietzel, P. D. C. (2019). Cryst. Growth Des. 19, 2089–2096.  CrossRef CAS Google Scholar
First citationShields, G. P., Raithby, P. R., Allen, F. H. & Motherwell, W. D. S. (2000). Acta Cryst. B56, 455–465.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationStreek, J. van de & Neumann, M. A. (2014). Acta Cryst. B70, 1020–1032.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSykes, R. A., McCabe, P., Allen, F. H., Battle, G. M., Bruno, I. J. & Wood, P. A. (2011). J. Appl. Cryst. 44, 882–886.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationToby, B. H. & Von Dreele, R. B. (2013). J. Appl. Cryst. 46, 544–549.  Web of Science CrossRef CAS IUCr Journals 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