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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

An ortho­rhom­bic polymorph of 2-(1,3,5-di­thia­zinan-5-yl)ethanol or MEA-di­thia­zine

crossmark logo

aDavidson School of Chemical Engineering, Purdue University, 480 W. Stadium Ave., West Lafayette, IN 47907, USA, bPioneer Oil, 400 Main Street, Vincennes, IN, 47951, USA, and cPurdue University, Department of Chemistry, 560 Oval Dr., West Lafayette, IN 47907, USA
*Correspondence e-mail: nschulth@purdue.edu

Edited by W. T. A. Harrison, University of Aberdeen, Scotland (Received 7 January 2022; accepted 10 January 2022; online 14 January 2022)

Substituted triazines are a class of compounds utilized for scavenging and sequestering hydrogen sulfide in oil and gas production operations. The reaction of one of these triazines under field conditions resulted in the formation of the title compound, 2-(1,3,5-di­thia­zinan-5-yl)ethanol, C5H11NOS2, or MEA-di­thia­zine. Polymorphic form I, in space group I41/a, was first reported in 2004 and its extended structure displays one-dimensional, helical strands connected through O—H⋯O hydrogen bonds. We describe here the form II polymorph of the title compound, which crystallizes in the ortho­rhom­bic space group Pbca as centrosymmetric dimers through pairwise O—H⋯N hydrogen bonds from the hydroxyl moiety to the nitro­gen atom of an adjacent mol­ecule.

1. Chemical context

Hydrogen sulfide is a corrosive and lethal gas that is commonly encountered during the production of hydro­carbons from subterranean reservoirs (Marriott et al., 2016[Marriott, R., Pirzadeh, P., Marrugo-Hernandez, J. & Raval, S. (2016). Can. J. Chem. 94, 406-413.]). The highly toxic H2S gas must be removed from the hydro­carbon stream to ensure a commercially salable product into the pipeline and refinery distribution network (Kermani et al., 2006[Kermani, B., Martin, J. & Esaklul, K. (2006). NACE Annual Corrosion Conference, Paper No. 06121.]). One of the most widely applied economical and effective strategies is to scavenge the hydrogen sulfide through the use of scavenger chemicals such as hexa­hydro-1,3,5-tris­(hy­droxy­eth­yl)-s-triazine, colloquially referred to as mono­ethano­lamine triazine (MEA-triazine) (Taylor et al., 2017[Taylor, G., Wylde, J., Müller, T., Murison, J. & Schneider, F. (2017). SPE-184529.]). The MEA-triazine is routinely administered at remote field locations using engineering controls that enhance the efficiency with which the hydrogen sulfide reacts with the MEA-triazine scavenger.

[Scheme 1]

The reaction of hydrogen sulfide with the MEA-triazine (1) scavenger has been described, see Fig. 1[link] (Bakke et al., 2001[Bakke, J., Buhaug, J. & Riha, J. (2001). Ind. Eng. Chem. Res. 40, 6051-6054.]; Wylde et al., 2020[Wylde, J., Taylor, G., Sorbie, K. & Samaniego, W. (2020). Energy Fuels, 34, 9923-9931.]). Furthermore, while hydrogen sulfide can theoretically react with a third mol­ecule to yield a tri­thiane mol­ecule, this reaction does not proceed under most conditions. As a result, sparingly soluble solid di­thia­zine (3) mol­ecules or an amorphous polymerized di­thia­zine material are the typical final products under aqueous reaction conditions (Taylor & Matherly, 2011[Taylor, G. & Matherly, R. (2011). Ind. Eng. Chem. Res. 50, 735-740.]; Taylor et al., 2013[Taylor, G., Prince, P., Matherly, R., Ponnapati, R., Tompkins, R., Panchalingam, V., Jovancicevic, V. & Ramachandran, S. (2013). SPE-164134.]; Wang et al., 2020[Wang, S., Madekufamba, M., Lesage, K. L., Bernard, F., Gelfand, B., Davis, P. M., Tittemore, K., Marriott, R. A. & Botros, K. K. (2020). J. Nat. Gas Sci. Eng. 78, 103286, 10 pages.]). Engineering protocols are implemented into the pipeline flow systems so that the sparingly soluble solids from the spent scavenger are continuously removed. However, these operations are occasionally not completely effective in eliminating solids and a buildup of intra­ctable residues in the system will result, thus impeding the flow path and complicating field operations. This was the case at one production facility, where the title compound (3) crystallized within the field treatment system.

[Figure 1]
Figure 1
Reaction of hydrogen sulfide with 2-(1,3,5-di­thia­zinan-5-yl)ethanol (MEA-triazine) (1) to yield the MEA mono- (2) and di­thia­zine (3) products.

The mol­ecular and crystal structure of title compound 3 has been determined before. The first polymorph, form I, in space group I41/a was described independently by Galvez-Ruiz et al. (2004[Galvez-Ruiz, J., Jaen-Gaspar, J., Castellanos-Arzola, I., Contreras, R. & Flores-Parra, A. (2004). Heterocycles, 63, 2269-2285.]) and Wang et al. (2020[Wang, S., Madekufamba, M., Lesage, K. L., Bernard, F., Gelfand, B., Davis, P. M., Tittemore, K., Marriott, R. A. & Botros, K. K. (2020). J. Nat. Gas Sci. Eng. 78, 103286, 10 pages.]). Recently, a second polymorph, in space group Pbca, was disclosed as a CSD communication (Unruh, 2021[Unruh, D. K. (2021). CSD Communication (refcode ACAMIB02, CCDC 2071625). CCDC, Cambridge, England.]). No further details, such as the origin of the material, the solvent or mode of crystallization or other information was stated with the deposited cif file and the mol­ecular and crystal structures were not discussed.

2. Structural commentary

Single crystals formed in the field treatment system and were used for intensity data collection without recrystallization. Data were collected at both room temperature as well as 150 K. The polymorph obtained was that of form II, in space group Pbca, and the data are in good agreement with those recently disclosed by Unruh (2021[Unruh, D. K. (2021). CSD Communication (refcode ACAMIB02, CCDC 2071625). CCDC, Cambridge, England.]).

Phase purity of the sample as harvested from the field treatment system was verified by powder X-ray diffraction analysis (supplementary Figures S1 and S2). No phase change was observed between 150 K and room temperature. The unit-cell parameters obtained by Rietveld analysis of the structural model against the room-temperature powder pattern are a = 9.6596 (6), b = 8.9248 (5), c = 17.748 (1) Å and V = 1530.03 (15) Å3 (Table S1A), matching the parameters from the room-temperature single-crystal data collection (Table 4[link]). Powder data and unit-cell parameters from samples recrystallized from either iso­propanol or an iso­propanol–water mixture are indistinguishable from those of the field sample (Fig. S1, Fig. S2, Table S1).

Table 4
Experimental details

  150 K RT
Crystal data
Chemical formula C5H11NOS2 C5H11NOS2
Mr 165.27 165.27
Crystal system, space group Orthorhombic, Pbca Orthorhombic, Pbca
Temperature (K) 150 296
a, b, c (Å) 9.5828 (4), 8.8467 (3), 17.6400 (8) 9.7160 (11), 8.9709 (10), 17.8577 (17)
V3) 1495.45 (11) 1556.5 (3)
Z 8 8
Radiation type Mo Kα Mo Kα
μ (mm−1) 0.63 0.61
Crystal size (mm) 0.41 × 0.34 × 0.28 0.41 × 0.34 × 0.28
 
Data collection
Diffractometer Bruker AXS D8 Quest diffractometer with PhotonII charge-integrating pixel array detector (CPAD) Bruker AXS D8 Quest diffractometer with PhotonII charge-integrating pixel array detector (CPAD)
Absorption correction Multi-scan (SADABS; Krause et al., 2015[Krause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3-10.]) Multi-scan (SADABS; Krause et al., 2015[Krause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3-10.])
Tmin, Tmax 0.689, 0.747 0.696, 0.747
No. of measured, independent and observed [I > 2σ(I)] reflections 60184, 2857, 2609 47635, 2909, 2378
Rint 0.040 0.040
(sin θ/λ)max−1) 0.770 0.765
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.018, 0.053, 1.06 0.027, 0.075, 1.02
No. of reflections 2857 2909
No. of parameters 86 86
H-atom treatment H atoms treated by a mixture of independent and constrained refinement H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.38, −0.21 0.33, −0.27
Computer programs: APEX3 and SAINT (Bruker, 2019[Bruker (2019). APEX3 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXS (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]), SHELXL2018/3 (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. C71, 3-8.]), shelXle (Hübschle et al., 2011[Hübschle, C. B., Sheldrick, G. M. & Dittrich, B. (2011). J. Appl. Cryst. 44, 1281-1284.]), 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.]), and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

The mol­ecular conformations and packing at low temperature and room temperature are essentially identical and unless otherwise noted, the following discussion will be based on the 150 K data set. The title compound (Fig. 2[link]) crystallizes in the ortho­rhom­bic space group Pbca. Bond distances and angles are in the expected ranges and unexceptional. Key torsion angles are given in Table 1[link]. The nitro­gen atom is sp3 hybridized, as expected for a tris­(methyl­ene) substituted amine [C—N—C angles range from 113.16 (5) to 114.49 (5) °, the bond angle sum is 342.06°], the di­thia­zine ring exhibits the expected chair conformation, and the ethano­lamine unit is substituted onto the ring in an axial orientation. The title compound has this in common with the other structures featuring a MEA-di­thia­zine moiety: The previously reported polymorph (form I) in space group I41/a and also the bromide and chloride salts of MEA-di­thia­zine, which are protonated at the amine and isomorphous (Bushmarinov et al., 2009[Bushmarinov, I., Antipin, M., Akhmetova, V., Nadyrgulova, G. & Lyssenko, K. (2009). Mendeleev Commun. 19, 14-16.]; Galvez-Ruiz et al., 2008[Galvez-Ruiz, J. C., Solano-Ruiz, E., Sanchez-Ruiz, S. A., Contreras, R. & Flores-Parra, A. (2008). Arkivoc, pp. 81-85.]), also feature an sp3-hybridized amine N atom, as well as a chair conformation di­thia­zine ring axially substituted by the ethano­lamine.

Table 1
Selected torsion angles (°) for the 150 K structure[link]

C2—S2—C1—S1 60.15 (4) C4—N1—C3—S1 67.01 (7)
C3—S1—C1—S2 −60.52 (4) C1—S1—C3—N1 60.35 (5)
C3—N1—C2—S2 65.52 (6) C2—N1—C4—C5 −104.74 (6)
C4—N1—C2—S2 −68.05 (6) C3—N1—C4—C5 122.32 (6)
C1—S2—C2—N1 −58.98 (5) N1—C4—C5—O1 65.09 (7)
C2—N1—C3—S1 −66.53 (6)    
[Figure 2]
Figure 2
The mol­ecular structure of 2-(1,3,5-di­thia­zinan-5-yl)ethanol. Displacement ellipsoids are shown at the 50% probability level.

The most flexible fragment of the MEA-di­thia­zine mol­ecule, and worth a closer investigation, is the ethano­lamine chain. In the form II polymorph the N1—C4—C5—O1 chain adopts a gauche conformation, with a torsion angle of 65.09 (7)° (Table 2[link]). In the form I polymorph the conformation is also gauche, with a torsion angle of −70.1 (3)° (in the chosen enanti­omer). In the bromide and chloride salts, this torsion angle is slightly smaller (55.1, 54.2 and −55.0° in the three published structures), but also gauche. The form I polymorph and the salts thus exhibit the same chair conformation and axial position of the ethano­lamine group as well as a common gauche conformation of the ethano­lamine chain as observed for form II of the title compound, indicating a clear preference of the MEA-di­thia­zine mol­ecule towards this specific mol­ecular conformation.

Table 2
Hydrogen-bond geometry (Å, °) for the 150 K[link] structure

D—H⋯A D—H H⋯A DA D—H⋯A
O1—H1⋯N1i 0.792 (13) 2.124 (13) 2.9103 (8) 172.3 (13)
C3—H3A⋯O1ii 0.99 2.42 3.3359 (9) 154
C2—H2A⋯S1iii 0.99 3.01 3.9330 (7) 155
C3—H3B⋯S2iv 0.99 2.96 3.7899 (7) 142
Symmetry codes: (i) [-x+1, -y, -z+1]; (ii) [x+{\script{1\over 2}}, -y+{\script{1\over 2}}, -z+1]; (iii) [-x+1, y-{\script{1\over 2}}, -z+{\script{3\over 2}}]; (iv) [x+{\script{1\over 2}}, y, -z+{\script{3\over 2}}].

What does differ between the three structures involving MEA-di­thia­zine or its cation are the C—N—C—C torsion angles between the di­thia­zine ring and the ethyl group. In form II, they are −104.74 (6)° (C2—N1—C4—C5) and 122.32 (6)° (C3—N1—C4—C5). The equivalent torsion angles in form I are 78.9 (3) and −148.1 (3)°, respectively, and those of the chloride salt are 175.8 and −55.2° (Bushmarinov et al., 2009[Bushmarinov, I., Antipin, M., Akhmetova, V., Nadyrgulova, G. & Lyssenko, K. (2009). Mendeleev Commun. 19, 14-16.]). Fig. 3[link], an overlay of the mol­ecules from the three crystalline motifs, exemplifies the similarities and differences. The cause for the differences between the two polymorphs and the salt can be found among the inter­molecular and packing inter­actions, to be discussed below.

[Figure 3]
Figure 3
An overlay of the two polymorphs of MEA-di­thia­zine (red: form II; blue form I) and of the cation of the chloride salt (green). Overlays are based on a least-squares fit of the di­thia­zine ring atoms.

3. Supra­molecular features

Different inter­molecular inter­actions, especially hydrogen bonds, are the core reason for the formation of the two MEA-di­thia­zine polymorphs (form I and II). In the previously described crystal structure (form I), the mol­ecules are connected through O—H⋯O hydrogen bonds involving the hydroxyl O atoms as both hydrogen-bond donor and acceptor. A series of these inter­actions lead to the formation of one-dimensional, helical spirals in which adjacent mol­ecules are related to each other via the 41 screw axis of the I41/a space group (Figs. 4[link] and 5[link]). The graph-set motif of the infinite strands is C11(2). The helical chains are further supported by weak C—H⋯S inter­actions involving the methyl­ene group adjacent to the OH group and one of the sulfur atoms of a mol­ecule related by three turns of the screw axis [C5⋯S2i = 3.533 (3) Å; symmetry code: (i) = −[{3\over 4}] + y, 3/4 – x, 3/4 – z; Fig. 5[link]]. No other significant inter­actions involving H atoms are observed, and individual helical spirals run parallel and anti-parallel to each other with no directional inter­actions between them. One other important aspect of the form-I polymorph is that the amine N atom does not act as a hydrogen-bond acceptor, neither towards the O—H group nor any of the methyl­ene CH2 hydrogen atoms.

[Figure 4]
Figure 4
Hydrogen-bonding inter­actions in the form I polymorph. Mol­ecules are linked into infinite chains that extend along the c-axis direction. Hydrogen-bonded mol­ecules are symmetry-related to each other via the 41-screw axis of the I41/a space group. Coordinates from Wang et al. (2020[Wang, S., Madekufamba, M., Lesage, K. L., Bernard, F., Gelfand, B., Davis, P. M., Tittemore, K., Marriott, R. A. & Botros, K. K. (2020). J. Nat. Gas Sci. Eng. 78, 103286, 10 pages.]). Displacement ellipsoids are shown at the 50% probability level. H atoms not involved in hydrogen-bonding are omitted for clarity.
[Figure 5]
Figure 5
Partial packing view and hydrogen bonding in the form I polymorph, viewed at a slight angle along the c axis (41-screw-axis direction). Turquoise dashed lines: O—H⋯O hydrogen bonds. Light-green dashed lines: weak C—H⋯S inter­actions. Coordinates from Wang et al. (2020[Wang, S., Madekufamba, M., Lesage, K. L., Bernard, F., Gelfand, B., Davis, P. M., Tittemore, K., Marriott, R. A. & Botros, K. K. (2020). J. Nat. Gas Sci. Eng. 78, 103286, 10 pages.]). Displacement ellipsoids are shown at the 50% probability level.

In the form II structure described here, the amine N atom behaves differently and acts as the primary hydrogen-bond acceptor (Tables 2[link] and 3[link]). Pairwise O—H⋯N [O⋯N = 2.9103 (8) Å] hydrogen bonds connect MEA-di­thia­zine mol­ecules into centrosymmetric dimers, with a graph-set notation of R22(10). (Table 3[link], Figs. 6[link] and 7[link]). No O—H⋯O inter­actions are observed in the form II polymorph.

Table 3
Hydrogen-bond geometry (Å, °) for the RT[link] structure

D—H⋯A D—H H⋯A DA D—H⋯A
O1—H1⋯N1i 0.79 (2) 2.18 (2) 2.9627 (12) 171 (2)
C3—H3A⋯O1ii 0.97 2.50 3.3979 (14) 153
C2—H2A⋯S1iii 0.97 3.10 3.9985 (11) 156
C3—H3B⋯S2iv 0.97 3.05 3.8512 (11) 141
Symmetry codes: (i) [-x+1, -y, -z+1]; (ii) [x+{\script{1\over 2}}, -y+{\script{1\over 2}}, -z+1]; (iii) [-x+1, y-{\script{1\over 2}}, -z+{\script{3\over 2}}]; (iv) [x+{\script{1\over 2}}, y, -z+{\script{3\over 2}}].
[Figure 6]
Figure 6
Hydrogen-bonding inter­actions in the form II polymorph. Mol­ecules are linked into centrosymmetric dimers connected via pairs of O—H⋯N hydrogen bonds. Displacement ellipsoids are shown at the 50% probability level.
[Figure 7]
Figure 7
Partial packing view and hydrogen bonding in the form II polymorph, viewed down the c axis. Turquoise dashed lines: O—H⋯N hydrogen bonds. Light-green dashed lines: C—H⋯O inter­actions. Displacement ellipsoids are shown at the 50% probability level.

Contrary to the form I polymorph, where the spiral chains propagate in the c-axis direction with no significant directional inter­actions between parallel strands, the primary building units in the form II structure are connected to each other via C—H⋯O hydrogen bonds and weak C—H⋯S inter­actions. The C—H⋯O inter­actions are notably short for this kind of hydrogen bond [H⋯O = 2.42 Å, C⋯O = 3.3359 (9)Å]. The C—H⋯O inter­actions connect the hydrogen-bonded dimers into infinite layers lying perpendicular to (001).

The C—H⋯S inter­action is much weaker and connects parallel layers with each other. The layers, which are slightly corrugated, also inter­digitate with each other, yielding a densely packed and rigid three-dimensional arrangement. This is reflected in the density of the crystals in form II, which is (at 150 K) 1.468 g cm−3. The form I structure, with its lack of strong inter­actions between hydrogen-bonded spirals, is substanti­ally less densely packed (1.407 g cm−3 at 173 K, or 4.3% less dense than form II). This points towards form II likely being the thermodynamically more stable polymorph. This is also supported by the melting temperatures of the two forms, where form I has a reported melting onset of 314 K in comparison to 316 K measured for form II (Figure S1, supporting information). The difference in melting point between the two forms is small (∼2K); this could be coincidental or they might point towards a phase transformation of one of the two forms upon heating, with only one form being present once the melting point temperature is reached. For form II, a differential scanning calorimetry (DSC) investigation did show any indication of a phase change (Figure S2, supporting information). However, a thorough investigation of both polymorphs utilizing DSC would be necessary to determine the relative stabilities (Yu, 1995[Yu, L. (1995). J. Pharm. Sci. 84, 966-974.]).

Crystals of the form I polymorph as reported by both Galvez-Ruiz et al. (2004[Galvez-Ruiz, J., Jaen-Gaspar, J., Castellanos-Arzola, I., Contreras, R. & Flores-Parra, A. (2004). Heterocycles, 63, 2269-2285.]) and Wang et al. (2020[Wang, S., Madekufamba, M., Lesage, K. L., Bernard, F., Gelfand, B., Davis, P. M., Tittemore, K., Marriott, R. A. & Botros, K. K. (2020). J. Nat. Gas Sci. Eng. 78, 103286, 10 pages.]) are not the original material as isolated from the reaction of hydrogen sulfide with the MEA-triazine 1 scavenger in water as the solvent. Galvez-Ruiz et al. analyzed laboratory-prepared material that was obtained via a different route (reaction of ethano­lamine with NaHS and formaldehyde) and crystals were grown from chloro­form solution. In the 2020 report, the original material had been field samples obtained from an unspecified natural gas site, but the samples were purified and recrystallized prior to analysis. Field samples containing >90 wt% di­thia­zine were dissolved in isopropyl alcohol, filtered, and di­thia­zine crystals were obtained from iso­propanol–water mixtures by cooling to 278 K, and recrystallized twice to obtain large translucent crystals suitable for single-crystal structure analysis. All further analysis, including the measurement of the solubility of di­thia­zine in natural gas mixtures and variable pressure, was performed on the recrystallized samples. The original material from the field site was not further analyzed to verify that recrystallized and raw material were of the same kind, i.e., no powder XRD data of the original material were recorded, and no Rietveld analysis of the pattern was performed. The substantial difference in structure and density observed for the two polymorphs does indicate that they would also differ in other physicochemical properties. A noteworthy difference described herein is that crystalline material directly from the production facility and from laboratory recrystallization experiments were analyzed, and both resulted in the form II polymorph.

4. Database survey

A search of the Cambridge Structural Database (CSD, version 5.42 of November 2020; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) found a variety of hits featuring the 2-(1,3,5-di­thia­zinan-5-yl)ethanol structural backbone. The form I (I41/a) polymorph of 2-(1,3,5-di­thia­zinan-5-yl)ethanol, collected at 223 K, was first reported in 2004 (CSD refcode ACAMIB; Galvez-Ruiz et al., 2004[Galvez-Ruiz, J., Jaen-Gaspar, J., Castellanos-Arzola, I., Contreras, R. & Flores-Parra, A. (2004). Heterocycles, 63, 2269-2285.]) using room-temperature data, while a 173 K dataset of the same polymorph was reported in 2020 (ACAMIB01; Wang et al., 2020[Wang, S., Madekufamba, M., Lesage, K. L., Bernard, F., Gelfand, B., Davis, P. M., Tittemore, K., Marriott, R. A. & Botros, K. K. (2020). J. Nat. Gas Sci. Eng. 78, 103286, 10 pages.]). Form II, discussed in detail here, was recently disclosed as a CSD communication (ACAMIB02, Unruh 2021[Unruh, D. K. (2021). CSD Communication (refcode ACAMIB02, CCDC 2071625). CCDC, Cambridge, England.]). When a phenyl group is added to the ethanol moiety, a discrete, monomeric unit results with the formation of an intra­molecular O—H⋯N (O⋯N = 2.782 Å) hydrogen bond (ACAMOH; Galvez-Ruiz et al., 2004[Galvez-Ruiz, J., Jaen-Gaspar, J., Castellanos-Arzola, I., Contreras, R. & Flores-Parra, A. (2004). Heterocycles, 63, 2269-2285.]). Similarly, when a methyl group is appended (ACAMUN; Galvez-Ruiz et al., 2004[Galvez-Ruiz, J., Jaen-Gaspar, J., Castellanos-Arzola, I., Contreras, R. & Flores-Parra, A. (2004). Heterocycles, 63, 2269-2285.]; ACAMUN01; Colorado-Peralta et al., 2010[Colorado-Peralta, R., Xotlanihua-Flores, A., Gálvez-Ruíz, J. C., Sánchez-Ruíz, S., Contreras, R. & Flores-Parra, A. (2010). J. Mol. Struct. 981, 21-33.]), discrete monomers result with intra­molecular O—H⋯N (O⋯N = 2.721 and 2.732 Å) hydrogen bonds. However, when a phenyl and a methyl group are appended to the ethanol moiety (ACANAU; Galvez-Ruiz et al., 2004[Galvez-Ruiz, J., Jaen-Gaspar, J., Castellanos-Arzola, I., Contreras, R. & Flores-Parra, A. (2004). Heterocycles, 63, 2269-2285.]), one-dimensional strands form through O—H⋯N (O⋯N = 3.145 Å) hydrogen bonds with a C22(10) graph-set motif. Three crystalline organic salts of 2-(1,3,5-di­thia­zinan-5-yl)ethanol have also been reported, including one with bromide and two with chloride counter-ions (HOSKIK and HOSKOQ; Bushmarinov et al., 2009[Bushmarinov, I., Antipin, M., Akhmetova, V., Nadyrgulova, G. & Lyssenko, K. (2009). Mendeleev Commun. 19, 14-16.]; HOSKOQ01; Galvez-Ruiz et al., 2008[Galvez-Ruiz, J. C., Solano-Ruiz, E., Sanchez-Ruiz, S. A., Contreras, R. & Flores-Parra, A. (2008). Arkivoc, pp. 81-85.]). All three salts are isostructural, possessing a protonated nitro­gen atom and charge-assisted O—H⋯Br/Cl and charge-assisted N—H⋯Br/Cl hydrogen bonds and R42(14) graph sets. In all nine structures, the di­thia­zine ring adopts a chair conformation.

5. Methods

Powder diffraction data were collected in focusing mode on a Panalytical Empyrean X-ray diffractometer equipped with Bragg–Brentano HD optics, a sealed-tube copper X-ray source (λ = 1.54178 Å), Soller slits on both the incident and receiving optics sides, and a PixCel3D Medipix detector. Samples were hand ground for 20 minutes using an agate mortar and pestle and packed in metal sample cups with a sample area 16 mm wide and 2 mm deep. 1/4° anti-scatter slits and 1/16° divergence slits as well as a 4 mm mask were chosen based on sample area and starting θ angle. Data were collected between 5 and 90° in 2θ using Data Collector software (PANalytical, 2019[PANalytical (2019). Data Collector. PANalytical BV, Almelo, The Netherlands.]). Rietveld refinements were performed against the models of the single-crystal-structure data sets using the HighScore software (PANalytical, 2018[PANalytical (2018). HighScore. PANalytical BV, Almelo, The Netherlands.]). Refinement of preferred orientation was included using a spherical harmonics model.

Differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA) data were measured using a TA Instruments Q20 DSC with heating rates of 5°C min−1. The sample was run in an alumina open pan from 25 to 100°C. The DSC and TGA traces are given in the Supporting Information (Fig. S3).

5.1. Synthesis and crystallization

The commercially supplied MEA-triazine 1 solution, 47% active components in aqueous methanol (Innospec, Inc.), is stored in an external 650 gal polyethyl­ene tank and is dosed into two inclined static mixers with a piston-style positive displacement pump. A combined dose of 16 gal d−1 is administered to the two static mixers. Each mixer contains bead media that create a tortuous pathway that facilitates efficient mixing and provides the contact time for the hydrogen sulfide gases to react. One inclined static mixer services the well gases removed from the wellhead recovery line and the second services the excess gas from the stock-tank vapor-recovery system. The gaseous mixture of natural gas and hydrogen sulfide pass through the static mixer that contains the liquid MEA-triazine 1 solution. The MEA-triazine 1 solution is administered at a rate to keep the hydrogen sulfide below a targeted threshold and to ensure the di­thia­zine product 3 is transported through the static mixer at a sufficient rate to alleviate any plugging from reactionary solids.

The crystalline solids created through this process were collected and characterized by single-crystal and powder X-ray analysis as well as differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA). The X-ray powder pattern was successfully indexed suggesting a single, pure bulk phase. Thermal analysis revealed negligible weight loss through 100°C by TGA, with a melt endotherm onset at 43°C (574 J g−1) and a peak temperature at 45°C.

Additionally, the crystalline material from the field equipment was recrystallized from aqueous iso­propanol. Approximately 100 mg of crystalline material was dissolved with gentle heating in approximately 10 ml of neat IPA or 50:50 IPA:water. The samples were allowed to cool to room temperature and slowly evaporate. After 24 h, a crystalline material resulted, which was filtered, dried and characterized by powder X-ray diffraction (Fig. S1, S2D–S2I, Tables S1B, S1C).

6. Refinement

Single-crystal X-ray diffraction data, data collection and structure refinement details are summarized in Table 4[link]. A common structural model was refined against the data collected at 150 K and at room temperature (no phase change was observed). The C-bound H atoms were positioned geometrically and constrained to ride on their parent atoms with C—H bond distances of 0.99 Å (at 150 K) and 0.97 Å (at room temperature). The positions of the hydroxyl H atoms were refined. Uiso(H) values were set to 1.2Ueq(C) or 1.5Ueq(O).

Supporting information


Computing details top

For both structures, data collection: APEX3 (Bruker, 2019); cell refinement: SAINT (Bruker, 2019); data reduction: SAINT (Bruker, 2019). Program(s) used to solve structure: SHELXS (Sheldrick, 2008) for 150K; isomorphous replacement for RT. For both structures, program(s) used to refine structure: SHELXL2018/3 (Sheldrick, 2015), shelXle (Hübschle et al., 2011); molecular graphics: Mercury (Macrae et al., 2020); software used to prepare material for publication: publCIF (Westrip, 2010).

2-(1,3,5-Dithiazinan-5-yl)ethanol (150K) top
Crystal data top
C5H11NOS2Dx = 1.468 Mg m3
Mr = 165.27Mo Kα radiation, λ = 0.71073 Å
Orthorhombic, PbcaCell parameters from 9803 reflections
a = 9.5828 (4) Åθ = 2.3–33.1°
b = 8.8467 (3) ŵ = 0.63 mm1
c = 17.6400 (8) ÅT = 150 K
V = 1495.45 (11) Å3Block, colourless
Z = 80.41 × 0.34 × 0.28 mm
F(000) = 704
Data collection top
Bruker AXS D8 Quest
diffractometer with PhotonII charge-integrating pixel array detector (CPAD)
2857 independent reflections
Radiation source: fine focus sealed tube X-ray source2609 reflections with I > 2σ(I)
Triumph curved graphite crystal monochromatorRint = 0.040
Detector resolution: 7.4074 pixels mm-1θmax = 33.2°, θmin = 3.1°
ω and phi scansh = 1414
Absorption correction: multi-scan
(SADABS; Krause et al., 2015)
k = 1313
Tmin = 0.689, Tmax = 0.747l = 2727
60184 measured reflections
Refinement top
Refinement on F2Secondary atom site location: difference Fourier map
Least-squares matrix: fullHydrogen site location: mixed
R[F2 > 2σ(F2)] = 0.018H atoms treated by a mixture of independent and constrained refinement
wR(F2) = 0.053 w = 1/[σ2(Fo2) + (0.0247P)2 + 0.3304P]
where P = (Fo2 + 2Fc2)/3
S = 1.06(Δ/σ)max = 0.002
2857 reflectionsΔρmax = 0.38 e Å3
86 parametersΔρmin = 0.21 e Å3
0 restraintsExtinction correction: SHELXL-2018/3 (Sheldrick, 2015), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
Primary atom site location: structure-invariant direct methodsExtinction coefficient: 0.0043 (8)
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
S10.49529 (2)0.46751 (2)0.67098 (2)0.01985 (5)
S20.24939 (2)0.25649 (2)0.68835 (2)0.01846 (5)
O10.35020 (6)0.04378 (6)0.45260 (3)0.02346 (11)
H10.4068 (14)0.0193 (14)0.4443 (7)0.035*
N10.46380 (6)0.20571 (6)0.58421 (3)0.01538 (10)
C10.38115 (7)0.36827 (8)0.73583 (4)0.01980 (12)
H1A0.4380740.3010550.7683470.024*
H1B0.3345620.4428530.7691540.024*
C20.36757 (7)0.13153 (7)0.63564 (4)0.01697 (11)
H2A0.4225660.0721170.6726470.020*
H2B0.3104230.0593140.6060760.020*
C30.56527 (7)0.30023 (8)0.62278 (4)0.01882 (12)
H3A0.6352430.3341320.5851540.023*
H3B0.6146750.2374700.6607110.023*
C40.39690 (8)0.27738 (7)0.51799 (4)0.01898 (12)
H4A0.4361050.3799190.5108160.023*
H4B0.2955990.2879030.5276600.023*
C50.41878 (7)0.18614 (8)0.44649 (4)0.01854 (12)
H5A0.3812270.2423630.4024610.022*
H5B0.5198670.1701010.4381740.022*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
S10.02283 (9)0.01648 (8)0.02024 (9)0.00517 (6)0.00081 (6)0.00304 (5)
S20.01611 (8)0.01798 (8)0.02130 (9)0.00069 (5)0.00337 (5)0.00081 (5)
O10.0202 (2)0.0179 (2)0.0323 (3)0.00111 (19)0.0025 (2)0.0054 (2)
N10.0183 (2)0.0142 (2)0.0136 (2)0.00098 (18)0.00090 (18)0.00015 (17)
C10.0236 (3)0.0202 (3)0.0156 (3)0.0008 (2)0.0019 (2)0.0020 (2)
C20.0196 (3)0.0126 (2)0.0187 (3)0.0001 (2)0.0002 (2)0.0012 (2)
C30.0150 (3)0.0210 (3)0.0204 (3)0.0004 (2)0.0006 (2)0.0017 (2)
C40.0271 (3)0.0151 (2)0.0147 (3)0.0028 (2)0.0025 (2)0.0009 (2)
C50.0187 (3)0.0225 (3)0.0144 (3)0.0012 (2)0.0011 (2)0.0004 (2)
Geometric parameters (Å, º) top
S1—C11.8100 (7)C1—H1B0.9900
S1—C31.8338 (7)C2—H2A0.9900
S2—C11.8093 (7)C2—H2B0.9900
S2—C21.8355 (7)C3—H3A0.9900
O1—C51.4247 (9)C3—H3B0.9900
O1—H10.792 (13)C4—C51.5120 (9)
N1—C21.4505 (8)C4—H4A0.9900
N1—C31.4517 (9)C4—H4B0.9900
N1—C41.4758 (8)C5—H5A0.9900
C1—H1A0.9900C5—H5B0.9900
C1—S1—C397.04 (3)N1—C3—S1115.93 (5)
C1—S2—C297.65 (3)N1—C3—H3A108.3
C5—O1—H1107.0 (9)S1—C3—H3A108.3
C2—N1—C3113.16 (5)N1—C3—H3B108.3
C2—N1—C4114.41 (5)S1—C3—H3B108.3
C3—N1—C4114.49 (5)H3A—C3—H3B107.4
S2—C1—S1113.22 (4)N1—C4—C5111.76 (5)
S2—C1—H1A108.9N1—C4—H4A109.3
S1—C1—H1A108.9C5—C4—H4A109.3
S2—C1—H1B108.9N1—C4—H4B109.3
S1—C1—H1B108.9C5—C4—H4B109.3
H1A—C1—H1B107.7H4A—C4—H4B107.9
N1—C2—S2115.88 (4)O1—C5—C4110.17 (5)
N1—C2—H2A108.3O1—C5—H5A109.6
S2—C2—H2A108.3C4—C5—H5A109.6
N1—C2—H2B108.3O1—C5—H5B109.6
S2—C2—H2B108.3C4—C5—H5B109.6
H2A—C2—H2B107.4H5A—C5—H5B108.1
C2—S2—C1—S160.15 (4)C4—N1—C3—S167.01 (7)
C3—S1—C1—S260.52 (4)C1—S1—C3—N160.35 (5)
C3—N1—C2—S265.52 (6)C2—N1—C4—C5104.74 (6)
C4—N1—C2—S268.05 (6)C3—N1—C4—C5122.32 (6)
C1—S2—C2—N158.98 (5)N1—C4—C5—O165.09 (7)
C2—N1—C3—S166.53 (6)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O1—H1···N1i0.792 (13)2.124 (13)2.9103 (8)172.3 (13)
C3—H3A···O1ii0.992.423.3359 (9)154
C2—H2A···S1iii0.993.013.9330 (7)155
C3—H3B···S2iv0.992.963.7899 (7)142
Symmetry codes: (i) x+1, y, z+1; (ii) x+1/2, y+1/2, z+1; (iii) x+1, y1/2, z+3/2; (iv) x+1/2, y, z+3/2.
2-(1,3,5-Dithiazinan-5-yl)ethanol (RT) top
Crystal data top
C5H11NOS2Dx = 1.411 Mg m3
Mr = 165.27Mo Kα radiation, λ = 0.71073 Å
Orthorhombic, PbcaCell parameters from 9354 reflections
a = 9.7160 (11) Åθ = 3.1–32.9°
b = 8.9709 (10) ŵ = 0.61 mm1
c = 17.8577 (17) ÅT = 296 K
V = 1556.5 (3) Å3Block, colourless
Z = 80.41 × 0.34 × 0.28 mm
F(000) = 704
Data collection top
Bruker AXS D8 Quest
diffractometer with PhotonII charge-integrating pixel array detector (CPAD)
2909 independent reflections
Radiation source: fine focus sealed tube X-ray source2378 reflections with I > 2σ(I)
Triumph curved graphite crystal monochromatorRint = 0.040
Detector resolution: 7.4074 pixels mm-1θmax = 33.0°, θmin = 3.3°
ω and phi scansh = 1414
Absorption correction: multi-scan
(SADABS; Krause et al., 2015)
k = 1313
Tmin = 0.696, Tmax = 0.747l = 2727
47635 measured reflections
Refinement top
Refinement on F2Secondary atom site location: difference Fourier map
Least-squares matrix: fullHydrogen site location: mixed
R[F2 > 2σ(F2)] = 0.027H atoms treated by a mixture of independent and constrained refinement
wR(F2) = 0.075 w = 1/[σ2(Fo2) + (0.0313P)2 + 0.3838P]
where P = (Fo2 + 2Fc2)/3
S = 1.02(Δ/σ)max = 0.001
2909 reflectionsΔρmax = 0.33 e Å3
86 parametersΔρmin = 0.27 e Å3
0 restraintsExtinction correction: SHELXL-2018/3 (Sheldrick, 2015), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
Primary atom site location: isomorphous structure methodsExtinction coefficient: 0.030 (2)
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.

Refinement. solved by isomorphous replacement from its 150 K structure

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
S10.49595 (3)0.46191 (3)0.67128 (2)0.04261 (9)
S20.25087 (3)0.25541 (3)0.68813 (2)0.03921 (8)
O10.35211 (10)0.04855 (10)0.45103 (6)0.0515 (2)
H10.408 (2)0.014 (2)0.4441 (10)0.077*
N10.46246 (8)0.20389 (9)0.58454 (4)0.03187 (16)
C10.38192 (12)0.36489 (12)0.73521 (5)0.0410 (2)
H1A0.3373140.4374070.7673760.049*
H1B0.4362450.2994900.7667940.049*
C20.36706 (11)0.13141 (10)0.63551 (5)0.03458 (19)
H2A0.4199860.0734590.6712020.042*
H2B0.3112910.0621120.6069070.042*
C30.56354 (10)0.29615 (12)0.62292 (6)0.0390 (2)
H3A0.6312880.3286850.5865230.047*
H3B0.6109540.2346890.6594080.047*
C40.39676 (13)0.27614 (11)0.51883 (5)0.0400 (2)
H4A0.4338470.3758370.5129230.048*
H4B0.2986220.2850120.5277390.048*
C50.42022 (11)0.18902 (13)0.44745 (5)0.0401 (2)
H5A0.3855300.2454130.4051200.048*
H5B0.5181190.1735400.4401560.048*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
S10.04861 (16)0.03562 (14)0.04361 (15)0.01122 (11)0.00100 (11)0.00706 (10)
S20.03421 (13)0.03855 (14)0.04485 (15)0.00120 (9)0.00745 (10)0.00121 (10)
O10.0443 (4)0.0425 (4)0.0676 (6)0.0036 (4)0.0086 (4)0.0147 (4)
N10.0372 (4)0.0300 (3)0.0284 (3)0.0038 (3)0.0015 (3)0.0002 (3)
C10.0503 (6)0.0417 (5)0.0311 (4)0.0001 (4)0.0031 (4)0.0045 (4)
C20.0406 (5)0.0254 (4)0.0377 (4)0.0000 (3)0.0008 (4)0.0022 (3)
C30.0306 (4)0.0447 (5)0.0416 (5)0.0010 (4)0.0017 (4)0.0030 (4)
C40.0572 (6)0.0324 (4)0.0303 (4)0.0064 (4)0.0054 (4)0.0014 (3)
C50.0394 (5)0.0518 (6)0.0291 (4)0.0008 (4)0.0026 (4)0.0003 (4)
Geometric parameters (Å, º) top
S1—C11.8134 (11)C1—H1B0.9700
S1—C31.8407 (11)C2—H2A0.9700
S2—C11.8145 (11)C2—H2B0.9700
S2—C21.8425 (10)C3—H3A0.9700
O1—C51.4248 (15)C3—H3B0.9700
O1—H10.79 (2)C4—C51.5125 (14)
N1—C21.4527 (13)C4—H4A0.9700
N1—C31.4558 (13)C4—H4B0.9700
N1—C41.4847 (12)C5—H5A0.9700
C1—H1A0.9700C5—H5B0.9700
C1—S1—C397.22 (5)N1—C3—S1116.07 (7)
C1—S2—C297.66 (5)N1—C3—H3A108.3
C5—O1—H1107.2 (15)S1—C3—H3A108.3
C2—N1—C3112.95 (8)N1—C3—H3B108.3
C2—N1—C4114.60 (8)S1—C3—H3B108.3
C3—N1—C4114.45 (8)H3A—C3—H3B107.4
S1—C1—S2113.38 (5)N1—C4—C5112.07 (8)
S1—C1—H1A108.9N1—C4—H4A109.2
S2—C1—H1A108.9C5—C4—H4A109.2
S1—C1—H1B108.9N1—C4—H4B109.2
S2—C1—H1B108.9C5—C4—H4B109.2
H1A—C1—H1B107.7H4A—C4—H4B107.9
N1—C2—S2116.12 (6)O1—C5—C4110.43 (9)
N1—C2—H2A108.3O1—C5—H5A109.6
S2—C2—H2A108.3C4—C5—H5A109.6
N1—C2—H2B108.3O1—C5—H5B109.6
S2—C2—H2B108.3C4—C5—H5B109.6
H2A—C2—H2B107.4H5A—C5—H5B108.1
C3—S1—C1—S260.08 (7)C4—N1—C3—S167.18 (10)
C2—S2—C1—S159.74 (7)C1—S1—C3—N160.12 (9)
C3—N1—C2—S265.59 (10)C2—N1—C4—C5106.30 (10)
C4—N1—C2—S267.90 (9)C3—N1—C4—C5120.91 (10)
C1—S2—C2—N159.00 (8)N1—C4—C5—O166.01 (12)
C2—N1—C3—S166.38 (10)
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O1—H1···N1i0.79 (2)2.18 (2)2.9627 (12)171 (2)
C3—H3A···O1ii0.972.503.3979 (14)153
C2—H2A···S1iii0.973.103.9985 (11)156
C3—H3B···S2iv0.973.053.8512 (11)141
Symmetry codes: (i) x+1, y, z+1; (ii) x+1/2, y+1/2, z+1; (iii) x+1, y1/2, z+3/2; (iv) x+1/2, y, z+3/2.
 

Acknowledgements

Pioneer Oil thanks Tim Hall for providing access to the location and awareness of the resulting crystalline materials. NS thanks Tim Manship and Davin Piercey for the DSC and TGA data collection.

Funding information

MZ thanks the National Science Foundation for funding for the single-crystal X-ray diffractometer through the Major Research Instrumentation Program under Grant No. CHE 1625543.

References

First citationBakke, J., Buhaug, J. & Riha, J. (2001). Ind. Eng. Chem. Res. 40, 6051–6054.  Web of Science CrossRef CAS Google Scholar
First citationBruker (2019). APEX3 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationBushmarinov, I., Antipin, M., Akhmetova, V., Nadyrgulova, G. & Lyssenko, K. (2009). Mendeleev Commun. 19, 14–16.  Web of Science CSD CrossRef CAS Google Scholar
First citationColorado-Peralta, R., Xotlanihua-Flores, A., Gálvez-Ruíz, J. C., Sánchez-Ruíz, S., Contreras, R. & Flores-Parra, A. (2010). J. Mol. Struct. 981, 21–33.  CAS Google Scholar
First citationGalvez-Ruiz, J., Jaen-Gaspar, J., Castellanos-Arzola, I., Contreras, R. & Flores-Parra, A. (2004). Heterocycles, 63, 2269–2285.  CAS Google Scholar
First citationGalvez-Ruiz, J. C., Solano-Ruiz, E., Sanchez-Ruiz, S. A., Contreras, R. & Flores-Parra, A. (2008). Arkivoc, pp. 81–85.  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 citationHübschle, C. B., Sheldrick, G. M. & Dittrich, B. (2011). J. Appl. Cryst. 44, 1281–1284.  Web of Science CrossRef IUCr Journals Google Scholar
First citationKermani, B., Martin, J. & Esaklul, K. (2006). NACE Annual Corrosion Conference, Paper No. 06121.  Google Scholar
First citationKrause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3–10.  Web of Science CSD CrossRef ICSD CAS IUCr Journals 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 citationMarriott, R., Pirzadeh, P., Marrugo-Hernandez, J. & Raval, S. (2016). Can. J. Chem. 94, 406–413.  Web of Science CrossRef CAS Google Scholar
First citationPANalytical (2018). HighScore. PANalytical BV, Almelo, The Netherlands.  Google Scholar
First citationPANalytical (2019). Data Collector. PANalytical BV, Almelo, The Netherlands.  Google Scholar
First citationSheldrick, G. M. (2008). Acta Cryst. A64, 112–122.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationTaylor, G. & Matherly, R. (2011). Ind. Eng. Chem. Res. 50, 735–740.  Web of Science CrossRef CAS Google Scholar
First citationTaylor, G., Prince, P., Matherly, R., Ponnapati, R., Tompkins, R., Panchalingam, V., Jovancicevic, V. & Ramachandran, S. (2013). SPE-164134.  Google Scholar
First citationTaylor, G., Wylde, J., Müller, T., Murison, J. & Schneider, F. (2017). SPE-184529.  Google Scholar
First citationUnruh, D. K. (2021). CSD Communication (refcode ACAMIB02, CCDC 2071625). CCDC, Cambridge, England.  Google Scholar
First citationWang, S., Madekufamba, M., Lesage, K. L., Bernard, F., Gelfand, B., Davis, P. M., Tittemore, K., Marriott, R. A. & Botros, K. K. (2020). J. Nat. Gas Sci. Eng. 78, 103286, 10 pages.  Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWylde, J., Taylor, G., Sorbie, K. & Samaniego, W. (2020). Energy Fuels, 34, 9923–9931.  Web of Science CrossRef CAS Google Scholar
First citationYu, L. (1995). J. Pharm. Sci. 84, 966–974.  CrossRef CAS PubMed Web of Science 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