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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Crystal structure of poly[μ-aqua-bis­(μ3-2-methylpropanoato-κ4O:O,O′:O′)dipotassium]

CROSSMARK_Color_square_no_text.svg

aInst. of Physics of the Czech Academy of Sciences, Na Slovance 2, 182 21 Praha 8, Czech Republic
*Correspondence e-mail: fabry@fzu.cz

Edited by G. Diaz de Delgado, Universidad de Los Andes, Venezuela (Received 15 May 2020; accepted 16 September 2020; online 22 September 2020)

The structure of the title compound, [K2(C4H7O2)2(H2O)]n, is composed of stacked sandwiches, which are formed by cation–oxygen bilayers surrounded by methyl­ethyl hydro­phobic chains. These sandwiches are held together by van der Waals inter­actions between the methyl­ethyl groups. The methyl­ethyl groups are disordered over two positions with occupancies 0.801 (3):0.199 (3). The potassium cations are coordinated by seven O atoms, which form an irregular polyhedron. There is a water mol­ecule, the oxygen atom of which is situated in a special position on a twofold axis (Wyckoff position 4e). The water H atoms are involved in Owater—H⋯Ocarbox­yl hydrogen bonds of moderate strength. These hydrogen bonds are situated within the cation–oxygen, i.e. hydro­philic, bilayer.

1. Chemical context

The structures of simple alkali 2-methyl­propano­ates (isobutyrates) have not been determined so far (as shown by a search of the Cambridge Structural Database, version 5.41, update of November 2019; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]) . In this context, `simple' means a compound containing just one cationic species. The reason for this rather surprising fact may follow from the expected difficult crystallization (Mirnaya et al., 1991[Mirnaya, T. A., Polishchuk, A. P., Molochaeva, V. I. & Tolochko, A. S. (1991). Kristallografiya, 36, 377-383.]). Moreover, the phases of isobutyrate salts are supposedly prone to undergo phase transitions due to the ordering of voluminous hydro­phobic methyl­ethyl chains by analogy to the phase transitions observed in alkali propionates and 2-methyl­propano­ates (Ferloni et al., 1975[Ferloni, P., Sanesi, M. & Franzosini, P. (1975). Z. Naturforsch. Teil A, 30, 1447-1457.]).

The chemistry of water solutions and the corresponding solid phases of 2-methyl­propano­ates and other carboxyl­ates where the number of carbon atoms is greater than two differs from that of formates and acetates. The structures and chemistry of the former compounds are also affected by hydration (Mirnaya et al., 1991[Mirnaya, T. A., Polishchuk, A. P., Molochaeva, V. I. & Tolochko, A. S. (1991). Kristallografiya, 36, 377-383.]). Hydration may take place because water mol­ecules compete with the carboxyl­ates in inclusion into the coordination sphere of the cations. Moreover, a rather tedious structure determination can be expected in alkanoates where the number of carbon atoms is greater than two because the organic chains tend to be positionally disordered and tend to exert large thermal agitation. This disorder, as well as the large displacement parameters, is related to the tendency to form different phases as pointed out above.

The structure determinations of 2-methyl­propano­ates as well as those of chemically related compounds with carboxyl­ates other than the formates and acetates show that their structures share the same tendency for the separation of metal cations, carboxyl­ate groups and sometimes water mol­ecules on the one hand from the organic chains on the other. The former groups are hydro­philic while the latter are hydro­phobic. The separation of these groups in these structures may be considered as an illustration of the alchemists' experience expressed by the slogan similis similibus solvuntur on a microstructural level. This separation also refers to solvate mol­ecules and affects their orientation with regard to their hydro­philic and hydro­phobic ends. The compound catena-[tetra­kis­(μ2-isobutyrato-O,O,O′)bis­(isobutyrato-O,O′)tri­aqua­dicerium ethanol solvate] (refcode XALZAN; Malaestean et al., 2012[Malaestean, I. L., Ellern, A., Baca, S. & Kögerler, P. (2012). Chem. Commun. 48, 1499-1501.]) can serve as an example.

Thus, the inter­molecular bonds in these structures can be divided into metal–oxygen bonds, O—H⋯O hydrogen bonds and van der Waals bonds between the hydro­phobic groups. The water mol­ecules as well as the solvate mol­ecules can be either coordinated to the cation or not while completing the hydro­philic part of the structures. At the same time, they are included into the hydrogen-bond pattern.

Correspondingly, the structures can be divided into the following classes (Table 1[link]):

Table 1
Overview of structural types observed in isobutyrates

Compound Refcode Reference
Water-free clusters
1 OHOXUF Malaestean et al. (2009[Malaestean, I. L., Speldrich, M., Ellern, A., Baca, S. G., Ward, M. & Kögerler, P. (2009). Eur. J. Inorg. Chem. pp. 4209-4212.])
2 GEWFUL Malaestean et al. (2013a[Malaestean, I. L., Ellern, A. & Kögerler, P. (2013a). Eur. J. Inorg. Chem., pp. 1635-1638.])
3 NAGQUI Coker et al. (2004[Coker, E. N., Boyle, T. J., Rodriguez, M. A. & Alam, T. M. (2004). Polyhedron, 23, 1739-1747.])
     
Clusters inter­connected by water mol­ecules
4 NAGQUI Coker et al. (2004[Coker, E. N., Boyle, T. J., Rodriguez, M. A. & Alam, T. M. (2004). Polyhedron, 23, 1739-1747.])
     
Water-free columns
5 PENJUN Ilina et al. (1992[Ilina, E. G., Troyanov, S. I. & Dunaeva, K. M. (1992). Koord. Khim. 18, 882-890.])
6 TAHXOR Boyle et al. (2010[Boyle, T. J., Raymond, R., Boye, D. M., Ottley, L. A. M. & Lu, P. (2010). Dalton Trans. 39, 8050-8063.])
7 TAHXOR01 Bierke & Meyer (2008[Bierke, T. & Meyer, G. (2008). Thesis, Universität zu Köln, Germany.])
8 TAHXOR02 Kotal et al. (2010[Kotal, A., Paira, T. K., Banerjee, S. & Mandal, T. K. (2010). Langmuir, 26, 6576-6582.])
     
Columns inter­connected by water mol­ecules
9 MECVAU Skelton & Deacon (2017[Skelton, B. W. & Deacon, G. B. (2017). CSD Communication (refcodes KELKOE, MECVAU). CCDC, Cambridge, England.])
10 XALZAN Malaestean et al. (2012[Malaestean, I. L., Ellern, A., Baca, S. & Kögerler, P. (2012). Chem. Commun. 48, 1499-1501.])
     
Water-free layered structures
11 KELKOE Skelton & Deacon (2017[Skelton, B. W. & Deacon, G. B. (2017). CSD Communication (refcodes KELKOE, MECVAU). CCDC, Cambridge, England.])
12 LUHGOK Yuranov & Dunaeva (1989[Yuranov, I. A. & Dunaeva, K. M. (1989). Koord. Khim. 15, 845-847.])
     
Structures with layers inter­connected by water mol­ecules
13 VIQTOG Malaestean et al. (2013b[Malaestean, I. L., Schmitz, S., Ellern, A. & Kögerler, P. (2013b). Acta Cryst. C69, 1144-1146.])
14 POSCIJ Troyanov et al. (1993[Troyanov, S. I., Mitrofanova, N. D. & Gorbacheva, M. V. (1993). Koord. Khim. 19, 868-870.])
15 SAJMUO Fischer et al. (2017[Fischer, A. I., Gurzhiy, V. V., Aleksandrova, J. V. & Pakina, M. I. (2017). Acta Cryst. E73, 318-321.])
Compound names, 1: bis(μ4-oxo)dodeca­kis­(μ3-isobutyrato)hexa­kis­(μ2-isobutyrato)bis(isobutyric acid)-bis­(propanol)-octa-manganese(II)-di-manganese(III) dihydrate; 2: hexa­kis­(μ3-isobutyrato)hexa­kis­(μ2-isobutyrato)hexa­kis­(2-methyl­propanoic acid)hexa­man­gan­ese; 3: hexa­kis­[bis­(μ2-2-methyl­propano­ato)(2-methyl­propanoic acid)magnesium]; 4: bis­(μ4-oxo)dodeca­kis­(μ3-isobutyrato)hexa­kis­(μ2-isobutyrato)bis­(isobutyric acid)bis­(propanol)octa­manganese(II)dimanganese(III) propanol solvate; 5: bis­[(μ3-isobutyrylato-O,O′)(μ2-isobutyrylato)copper(II)]; 6: catena-[tetra­kis­(μ2-2-methyl­prop­ano­ato)dizinc]; 7: catena-[tetra­kis­(μ2-2-methyl­propano­ato)dizinc]; 8: catena-[tetra­kis­(μ2-2-methyl­propano­ato)dizinc); 9: catena-[tetra­kis­(μ-2-methyl­propano­ato)bis­(2-methyl­propano­ato)tri­aqua­dilanthanum(III) hydrate; 10: catena-[tetra­kis­(μ2-isobutyrato-O,O,O′)bis­(isobutyrato-O,O′)tri­aqua­dicerium ethanol solvate]; 11: catena-[hexa­kis­(μ-methyl­propano­ato)dilutetium]; 12: di­aqua­bis­(isobutyrato)dioxouranium(VI); 13: catena-[(μ2-2-methyl­propano­ato)(2-methyl­propano­ato)tri­aqua­magnesium monohydrate]; 14: catena-[hexa­kis­(μ2-isobutyrato)aqua­dierbium monohydrate]; 15: catena-[bis­(μ-2-meth­yl­propano­ato)(μ-aqua)­cobalt(II) monohydrate].

(i) Structures that are composed of clusters where the inner part is formed by hydro­philic parts while the outer skin is formed by hydro­phobic groups.

(ii) Structures that are formed by columns, the inter­ior of which is composed of the hydro­philic parts while the outer skin is hydro­phobic.

(iii) Layered structures that are composed of stacked sandwiches formed by cation–oxygen bilayers surrounded by hydro­phobic organic groups. These sandwiches are bonded by van der Waals forces.

In all of these structural types, water mol­ecules can occur; examples are given in Table 1[link].

So far, the structures of 2-methyl­propano­ates (isobutyrates) have been reviewed. The motif of stacked layers, however, seems to be typical for simple alkali alkanoates M+CnH2n+1COO, n > 2, as follows from the known structures of Li(C3H5O2) (refcodes OMERUV, OMERUV01 and OMERUV02; Martínez Casado et al., 2009[Martínez Casado, F. J., Riesco, M. R., García Pérez, M. V., Redondo, M. I., López-Andrés, S. & Rodríguez Cheda, J. A. (2009). J. Phys. Chem. B, 113, 12896-12902.]), and the recently determined series of structures of Na, K, Rb and Cs propano­ates (Fábry & Samolová, 2020[Fábry, J. & Samolová, E. (2020). Acta Cryst. E76, 1508-1513.]), Tl(C3H5O2), catena-[(μ2-propano­ato)thallium(I)(propano­ato)thallium(I)] (WEWKAM; Martínez Casado et al., 2010[Martínez Casado, F. J., Ramos Riesco, M., da Silva, I., Labrador, A., Redondo, M. I., García Pérez, M. V., López-Andrés, S. & Rodríguez Cheda, J. A. (2010). J. Phys. Chem. B114, 10075-10085.]) and further from the structures of potassium acrylate and potassium methacrylate (refcodes VOVWOV and VOVWAH, respectively; Heyman et al., 2020[Heyman, J. B., Chen, C.-H. & Foxman, B. M. (2020). Cryst. Growth Des. 20, 330-336.]) as well as from the known structures with alkanoates with longer organic chains, e.g. potassium palmitate KC16H31O2 (KPALMA; Dumbleton & Lomer, 1965[Dumbleton, J. H. & Lomer, T. R. (1965). Acta Cryst. 19, 301-307.]).

Thus, the typical motif of separated hydro­phobic and hydro­philic parts of the mol­ecules can be generalized for carboxyl­ates other than formates and acetates.

The physical properties of 2-methyl­propano­ates as well as other related carboxyl­ates CnH2n+1COO, n > 2, hinder possible applications of these compounds, although there are some exceptions such as lanthanide zinc butyrates or their analogues, which have been applied for the synthesis of lanthanide–zinc–oxygen nanoparticles (Boyle et al., 2010[Boyle, T. J., Raymond, R., Boye, D. M., Ottley, L. A. M. & Lu, P. (2010). Dalton Trans. 39, 8050-8063.]) or for gelation induced by ultrasound in presence of ZnO nanoparticles (Kotal et al., 2010[Kotal, A., Paira, T. K., Banerjee, S. & Mandal, T. K. (2010). Langmuir, 26, 6576-6582.]).

[Scheme 1]

The aim of the present study was the preparation of potassium 2-methyl­propano­ate in order to fill the gap in the knowledge of these structures. Moreover, it was even more attractive to compare the sodium and potassium propano­ate structures (Fábry & Samolová, 2020[Fábry, J. & Samolová, E. (2020). Acta Cryst. E76, 1508-1513.]) in which the methyl groups are situated in two positions related by rotation of 180° because such a positional disorder mimics the arrangement of both methyl­ethyl chains in 2-methyl­propano­ates by the demand for space. However, a crystal of a hydrated phase has been obtained, the structure of which is reported here. Still, the authors believe that this reported structure determination adds a piece of knowledge that could be helpful in understanding the structural features in simple alkali carboxyl­ates with CnH2n+1COO, n > 2, and related structures.

2. Structural commentary

The structural unit of the title compound is shown in Fig. 1[link], which shows that the central cation is surrounded by seven oxygen atoms up to ∼3.33 Å. All of the oxygens stem from the carboxyl­ates except for the atom O3, which is a part of the coordinated water mol­ecule. The K—O bond distances are listed in Table 2[link]. Five of them, i.e. O2, O2ii, O2iii, O3 and O3iv, form a tetra­gonal pyramid with O2 as its apex. Atoms O1 and O1i complete the coordination polyhedron [symmetry codes: (i) 1 − x, y, [{1\over 2}] − z; (ii) [{3\over 2}] − x, −[{1\over 2}] + y, [{1\over 2}] − z; (iii) [{3\over 2}] − x, [{1\over 2}] + y, [{1\over 2}] − z; (iv) x, −1 + y, z.) It is also worth mentioning that the distances between the cation and the oxygen atoms belonging to the same carboxyl­ate are quite different: K1—O1 = 3.1113 (13) Å and K1—O2 = 2.8056 (13) Å.

Table 2
Selected bond lengths (Å)

K1—O1 3.1113 (13) K1—O2iii 2.7330 (12)
K1—O1i 2.6951 (14) K1—O3iv 3.3351 (13)
K1—O2 2.8056 (13) K1—O3 2.7693 (12)
K1—O2ii 2.7360 (12)    
Symmetry codes: (i) [-x+1, y, -z+{\script{1\over 2}}]; (ii) [-x+{\script{3\over 2}}, y-{\script{1\over 2}}, -z+{\script{1\over 2}}]; (iii) [-x+{\script{3\over 2}}, y+{\script{1\over 2}}, -z+{\script{1\over 2}}]; (iv) x, y-1, z.
[Figure 1]
Figure 1
A view of the structural motif in the title compound (DIAMOND; Brandenburg, 2005[Brandenburg, K. (2005). DIAMOND. Crystal Impact GbR, Bonn, Germany.]): Displacement ellipsoids are shown at the 30% probability level. K, O, C and water H atoms are shown as green, red, black ellipsoids as well as gray spheres, respectively. Symmetry codes: (i) 1 − x, y, [{1\over 2}] − z; (ii) [{3\over 2}] − x, −[{1\over 2}] + y, [{1\over 2}] − z; (iii) [{3\over 2}] − x, [{1\over 2}] + y, [{1\over 2}] − z; (iv) x, −1 + y, z.

3. Supra­molecular features

The prominent feature of the title structure is the presence of an oxygen–metal bilayer, which is surrounded by methyl­ethyl chains on both sides (Fig. 2[link]). This bilayer is composed of the cations and the oxygen atoms.

[Figure 2]
Figure 2
The packing of the mol­ecules in the title compound (DIAMOND; Brandenburg, 2005[Brandenburg, K. (2005). DIAMOND. Crystal Impact GbR, Bonn, Germany.]) viewed along the b axis. Displacement ellipsoids are shown at the 30% probability level. The colours are assigned to the atoms are as in Fig. 1[link].

Table 3[link] lists a pair of symmetry-equivalent Owater—H⋯Ocarboxyl­ate hydrogen bonds of moderate strength (Gilli & Gilli, 2009[Gilli, G. & Gilli, P. (2009). The Nature of the Hydrogen Bond: Outline of a comprehensive hydrogen bond theory, p. 61. Oxford University Press.]). These hydrogen bonds take place within the cation–oxygen bilayer (Fig. 2[link]). Inter­estingly, the water hydrogen atoms, supposedly positively charged, are directed towards the more distant cation K1 [H1O3⋯K1v and H1O3i⋯K1v = 3.033 (19) and 3.01 (2) Å, respectively, see Fig. 3[link] and its caption]. This means that the positive-charge inter­action diminishes a cohesive weak inter­action O3⋯K1v, the bond valence of which is 0.0385 (1) (Brese & O'Keeffe, 1991[Brese, N. E. & O'Keeffe, M. (1991). Acta Cryst. B47, 192-197.]). Other cohesive hydrogen-bonding inter­actions are listed in Table 3[link].

Table 3
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
O3—H1O3⋯O1v 0.83 (2) 1.92 (2) 2.7358 (17) 167 (2)
O3—H1O3i⋯O1vi 0.83 (2) 1.92 (2) 2.7358 (17) 167 (2)
Symmetry codes: (i) [-x+1, y, -z+{\script{1\over 2}}]; (v) x, y+1, z; (vi) [-x+1, y+1, -z+{\script{1\over 2}}].
[Figure 3]
Figure 3
The structural motif showing the inter­action of a water mol­ecule to K1v. Displacement ellipsoids are shown at the 30% probability level. The colours are assigned to the atoms are as in Fig. 1[link]. Symmetry codes as in Fig. 1[link] and (v) x, 1 + y, z.

As stated above, methyl­ethyl chains surround the hydro­philic inner bilayer on both sides. The packing of these sandwiches forms the title structure. The sandwiches are held together by van der Waals forces. Table 4[link] lists these weak inter­actions. Their distances are about the same as those in dicalcium barium hexa­kis­(propano­ate) Ca2Ba(C3H5O2)6 [4.05 (2) Å; Stadnicka & Glazer, 1980[Stadnicka, K. & Glazer, A. M. (1980). Acta Cryst. B36, 2977-2985.]] where disorder of the ethyl groups occurs. On the other hand, these inter­molecular distances are somewhat longer than in sodium and potassium propano­ates, where disorder of the ethyl groups has also been observed (Fábry & Samolová, 2020[Fábry, J. & Samolová, E. (2020). Acta Cryst. E76, 1508-1513.]). In the latter structures, the distances between two ethyl groups while one of them is in a disordered position are as short as 2.609 (8) and 2.651 (9) Å, respectively, which is an indication of a dynamic disorder: cf. the discussion about the disorder in Ca2Ba(C3H5O2)6 by Stadnicka & Glazer (1980[Stadnicka, K. & Glazer, A. M. (1980). Acta Cryst. B36, 2977-2985.]) according to whom the disorder is related to close C—C distances that are shorter than the sum of the van der Waals radii (about 4.5 Å). In the structurally related rubidium and caesium propano­ates, however, such an occupational disorder does not take place, most probably because of the longer distances between the ethyl groups in the latter structures. The shortest C—C distances in rubidium and caesium propano­ates are 3.908 (12) and 3.882 (13) Å, respectively.

Table 4
Cmethyl­ene—Cmeth­yl and Cmeth­yl—Cmeth­yl inter­molecular distances (Å) in the title structure of up to 4.5 Å

C2, C2a correspond to the methyl­ene carbon atoms while C3, C3a and C4, C4a correspond to methyl atoms.

C2⋯C4v 4.251 (6) C3⋯C4vi 3.992 (6)
C3⋯C4viii 4.136 (6) C4⋯C4ix 3.984 (5)
C3a⋯C4avii 4.17 (2) C3a⋯C4avi 4.25 (2)
C4a⋯C4aix 4.31 (2) C2a⋯C3ax 3.97 (2)
C2a⋯C4ax 3.884 (19) C2⋯C2av 3.960 (10)
C2⋯C4ax 4.479 (16) C3⋯C2av 4.128 (11)
C3⋯C3ax 4.44 (2) C3⋯C4avii 4.000 (15)
C3⋯C4avi 4.207 (15) C4⋯C3ax 4.48 (2)
C4⋯C3axi 4.087 (14) C4⋯C3aviii 4.43 (3)
C4⋯C4ax 3.767 (17) C4⋯C4aix 4.051 (17)
Symmetry codes: (v) x, y + 1, z; (vi) x − [{1\over 2}], y + [{1\over 2}], z; (vii) x − [{1\over 2}], y − [{1\over 2}], z; (viii) −x + 1, −y, −z; (ix) −x + [{3\over 2}], −y + [{1\over 2}], −z; (x) x, y − 1, z; (xi) x + [{1\over 2}], y − [{1\over 2}], z; (xii) x + [{1\over 2}], y + [{1\over 2}], z.

4. Synthesis and crystallization

Preparation of potassium 2-methyl­propano­ate was intended. The compound was prepared by dissolving potassium carbonate sesquihydrate (1.50 g) with 2-methyl­propanoic acid (0.80 g) in the molar ratio 1:2 in water. The pH of the solution was adjusted to 6–7 by addition of several tenths of a ml of the acid.

The solution was filtered and the excess amount of water was evaporated at 313 K. Shortly before crystallization, a layer with a pronounced viscosity appeared on the surface of the solution. The crystals grew in the form of elongated colourless plates of several tenths of a mm in their longest direction.

5. Structure determination and refinement

Crystal data, data collection and structure refinement details are summarized in Table 5[link]. The refinement was carried out on the averaged set of independent diffractions. All of the non-hydrogen atoms were determined by SHELXT (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. A71, 3-8.]). The structure was treated with consideration of a positional disorder of the methyl­ethyl chain. This disorder was revealed by a relatively high peak of residual electron density (0.68 e Å−3) in the vicinity of atom C2 (Fig. 4[link]a), which was pertinent to a model without assumed disorder. This residual peak was on the opposite side of the vector C2—H1C2 and was observable in the difference electron-density map using a model without the atoms C2, C3 and C4 as well as without the hydrogens attached to the latter carbons. This peak was assigned to a disordered atom C2 and denoted as C2a. Correspondingly, the carbon C3 was also disordered in the difference electron-density map (Fig. 4[link]b). Atoms C3 and C4 were split into the positions C3, C3a and C4, C4a after inclusion into the difference electron-density map.

Table 5
Experimental details

Crystal data
Chemical formula [K2(C4H7O2)2(H2O)]
Mr 270.4
Crystal system, space group Monoclinic, C2/c
Temperature (K) 240
a, b, c (Å) 11.9190 (5), 4.5454 (2), 24.3172 (9)
β (°) 97.517 (1)
V3) 1306.10 (9)
Z 4
Radiation type Cu Kα
μ (mm−1) 6.45
Crystal size (mm) 0.34 × 0.14 × 0.04
 
Data collection
Diffractometer Bruker D8 VENTURE Kappa Duo PHOTON 100 CMOS
Absorption correction Multi-scan (SADABS; Bruker, 2017[Bruker (2017). Instrument Service, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.])
Tmin, Tmax 0.219, 0.765
No. of measured, independent and observed [I > 3σ(I)] reflections 10532, 1269, 1212
Rint 0.036
(sin θ/λ)max−1) 0.618
 
Refinement
R[F > 3σ(F)], wR(F), S 0.028, 0.082, 2.48
No. of reflections 1269
No. of parameters 82
No. of restraints 9
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.19, −0.22
Computer programs: Instrument Service and SAINT (Bruker, 2017[Bruker (2017). Instrument Service, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXT (Sheldrick, 2015[Sheldrick, G. M. (2015). Acta Cryst. A71, 3-8.]), JANA2006 (Petříček et al., 2014[Petříček, V., Dušek, M. & Palatinus, L. (2014). Z. Kristallogr. 229, 345-352.]) and DIAMOND (Brandenburg, 2015).
[Figure 4]
Figure 4
(a) The maximum (0.6025, 0.0146, 0.1014), which is indicated by the arrow, is in the vicinity of C2. The increment of positive (solid lines) and negative (dashed) contours are 0.1 e Å−3. The height of the indicated maximum is 0.33 e Å−3 in the depicted section. The structural model did not contain the atoms C2, C2a, C3, C3a, C4 and C4a and the attached H atoms. (b) The maximum (0.4677, 0.5409, 0.0940), which is indicated by the arrow, is in the vicinity of C3. The increment of positive (solid lines) and negative (dashed) contours are 0.1 e Å−3. The height of the maximum is 0.27 e Å−3 in the depicted section. The model did not contain the atoms C2, C2a, C3, C3a, C4 and C4a and the attached H atoms.

The occupational parameters of these pairs of atoms, as well as of the attached hydrogens, were constrained so that their sum is equal to 1; the occupational parameter of atom C4 was refined. Each pair of these carbon atoms was constrained in such a way that the atom with the minor occupancy was assigned the same displacement parameters as the atom with the major occupancy. The carboxyl­ate carbon C1 was not split; the present model with the non-split carboxyl­ate carbon C1 was given preference because a splitting was too small and called for severe restraints of the C1—O1 and C1—O2 distances.

The methane­triyl hydrogens H1c2 and H1c2a, although observable, were placed in calculated positions. The latter hydrogens were refined under the following constraints: Cmethane­tri­yl—Hmethane­tri­yl = 0.99 Å, Uiso(Hmethane­tri­yl) = 1.2Ueq(Cmethane­tri­yl). Subsequently, after the anisotropic refinement of the non-hydrogen atoms with the methane­triyl hydrogen, the difference electron-density map revealed the methyl hydrogens. All of the methyl hydrogens were discernible in the difference electron-density maps. The hydrogens belonging to the major disorder component were found at first and then, after the refinement had converged, the other methyl hydrogens were found and refined. The methyl hydrogens were refined under the following constraints: Cmeth­yl—Hmeth­yl = 0.96 Å, Uiso(Hmethyl) = 1.5Ueq(Cmethyl). A following difference electron-density map revealed the water hydrogen, which was situated in a general position in contrast to its carrier O3. The water hydrogen was refined using the angle restraint H1O3—O3—H1O3i [symmetry code: (i) 1 − x, y, [{1\over 2}] − z] = 105.00 (1)° while Uiso(H1O3) = 1.5Ueq(O3). A trial refinement showed that the water oxygen was fully occupied. The C1—C2, C1—C2a bonds were restrained to be equal [1.540 (1) Å] as were C2—C3, C2—C3a and C2—C4, C2—C4a [1.500 (1) Å]. These values were found to yield the lowest R factors. Moreover, angle restraints to C3—C2—C4 and C3a—C2a—C4a were also applied. Of course, these C—C distances are affected by a large thermal agitation and are less reliable, as are the geometric parameters, compared to those of atom C1.

Supporting information


Computing details top

Data collection: Instrument Service (Bruker, 2017); cell refinement: SAINT (Bruker, 2017); data reduction: SAINT (Bruker, 2017); program(s) used to solve structure: SHELXT (Sheldrick, 2015); program(s) used to refine structure: JANA2006 (Petříček et al., 2014); molecular graphics: DIAMOND (Brandenburg, 2015); software used to prepare material for publication: JANA2006 (Petříček et al., 2014).

Poly[µ-aqua-bis(µ3-2-methylpropanoato-κ4O:O</i),O':O')dipotassium] top
Crystal data top
[K2(C4H7O2)2(H2O)]F(000) = 568
Mr = 270.4There have been used diffractions with I/σ(I)>20 for the unit cell determination.
Monoclinic, C2/cDx = 1.375 Mg m3
Hall symbol: -C 2ycCu Kα radiation, λ = 1.54178 Å
a = 11.9190 (5) ÅCell parameters from 8438 reflections
b = 4.5454 (2) Åθ = 7.4–72.2°
c = 24.3172 (9) ŵ = 6.45 mm1
β = 97.517 (1)°T = 240 K
V = 1306.10 (9) Å3Plate, colourless
Z = 40.34 × 0.14 × 0.04 mm
Data collection top
Bruker D8 VENTURE Kappa Duo PHOTON 100 CMOS
diffractometer
1269 independent reflections
Radiation source: IµS micro-focus sealed tube1212 reflections with I > 3σ(I)
Helios Cu multilayer optic monochromatorRint = 0.036
φ and ω scansθmax = 72.2°, θmin = 7.4°
Absorption correction: multi-scan
(SADABS; Bruker, 2017)
h = 1414
Tmin = 0.219, Tmax = 0.765k = 55
10532 measured reflectionsl = 3030
Refinement top
Refinement on F294 constraints
R[F > 3σ(F)] = 0.028Primary atom site location: dual
wR(F) = 0.082H atoms treated by a mixture of independent and constrained refinement
S = 2.48Weighting scheme based on measured s.u.'s w = 1/(σ2(I) + 0.0004I2)
1269 reflections(Δ/σ)max = 0.035
82 parametersΔρmax = 0.19 e Å3
9 restraintsΔρmin = 0.22 e Å3
Special details top

Refinement. The reflections 2 2 0, 4 2 0, 5 1 7 and 3 3 0 were excluded from the refinement because |Iobs-Icalc|>15σ(Iobs).

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/UeqOcc. (<1)
K10.65739 (3)0.08834 (7)0.292673 (13)0.03769 (13)
O10.54092 (11)0.0749 (2)0.17459 (5)0.0474 (4)
O20.71739 (10)0.0879 (2)0.18489 (5)0.0451 (4)
O30.50.5048 (4)0.250.0442 (5)
H1o30.521 (2)0.616 (3)0.2262 (8)0.0664*
C10.62018 (14)0.0693 (3)0.15869 (6)0.0357 (4)
C20.59779 (18)0.2545 (5)0.10542 (7)0.0445 (6)0.801 (3)
H1c20.620150.4586740.1159690.0533*0.801 (3)
C2a0.5957 (9)0.124 (2)0.09573 (13)0.0445 (6)0.199 (3)
H1c2a0.5908880.0556880.0726130.0533*0.199 (3)
C30.4752 (2)0.2726 (12)0.0811 (2)0.0828 (14)0.801 (3)
H1c30.4338610.3803140.1058860.1242*0.801 (3)
H2c30.4688920.3713380.0459750.1242*0.801 (3)
H3c30.4445360.0776880.0760860.1242*0.801 (3)
C3a0.4876 (10)0.297 (4)0.0912 (13)0.0828 (14)0.199 (3)
H1c3a0.4427740.2549280.0563680.1242*0.199 (3)
H2c3a0.446130.2433320.1209940.1242*0.199 (3)
H3c3a0.5046720.503460.0934010.1242*0.199 (3)
C40.6678 (4)0.1371 (12)0.06307 (12)0.1008 (19)0.801 (3)
H1c40.7438840.1038880.0803150.1512*0.801 (3)
H2c40.6358910.0448330.0483020.1512*0.801 (3)
H3c40.6682590.2773380.0335790.1512*0.801 (3)
C4a0.6882 (11)0.314 (4)0.0786 (7)0.1008 (19)0.199 (3)
H1c4a0.6697330.3679540.0403360.1512*0.199 (3)
H2c4a0.6958380.4878040.1011140.1512*0.199 (3)
H3c4a0.7581650.206390.0833890.1512*0.199 (3)
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
K10.0335 (2)0.0384 (2)0.0413 (2)0.00132 (11)0.00537 (14)0.00039 (11)
O10.0463 (7)0.0488 (7)0.0492 (7)0.0052 (5)0.0138 (5)0.0043 (5)
O20.0387 (6)0.0496 (6)0.0457 (7)0.0038 (5)0.0002 (5)0.0011 (4)
O30.0510 (10)0.0318 (7)0.0490 (9)00.0031 (7)0
C10.0375 (8)0.0359 (7)0.0345 (8)0.0047 (6)0.0075 (6)0.0002 (5)
C20.0494 (10)0.0422 (12)0.0417 (10)0.0012 (11)0.0056 (8)0.0081 (9)
C2a0.0494 (10)0.0422 (12)0.0417 (10)0.0012 (11)0.0056 (8)0.0081 (9)
C30.0537 (14)0.127 (3)0.061 (3)0.0035 (16)0.0173 (15)0.0378 (19)
C3a0.0537 (14)0.127 (3)0.061 (3)0.0035 (16)0.0173 (15)0.0378 (19)
C40.109 (3)0.155 (5)0.0460 (19)0.053 (3)0.038 (2)0.029 (2)
C4a0.109 (3)0.155 (5)0.0460 (19)0.053 (3)0.038 (2)0.029 (2)
Geometric parameters (Å, º) top
K1—O13.1113 (13)C2—C41.505 (4)
K1—O1i2.6951 (14)C2a—H1c2a0.99
K1—O22.8056 (13)C2a—C4a1.500 (18)
K1—O2ii2.7360 (12)C3—H1c30.96
K1—O2iii2.7330 (12)C3—H2c30.96
K1—O3iv3.3351 (13)C3—H3c30.96
K1—O32.7693 (12)C3a—H1c3a0.96
O1—C11.252 (2)C3a—H2c3a0.96
O2—C11.2499 (19)C3a—H3c3a0.96
O3—H1o30.83 (2)C4—H1c40.96
O3—H1o3i0.83 (2)C4—H2c40.96
C1—C21.539 (2)C4—H3c40.96
C1—C2a1.541 (4)C4a—H1c4a0.96
C2—C31.505 (4)C4a—H2c4a0.96
C2—C3a1.328 (13)C4a—H3c4a0.96
O1—K1—O1i84.51 (4)O2—C1—C2a122.5 (4)
O1—K1—O243.54 (3)C1—C2—H1c2106.47
O1—K1—O2ii98.72 (3)C1—C2—C3114.4 (2)
O1—K1—O2iii123.27 (4)C1—C2—C4109.3 (2)
O1—K1—O3iv50.08 (2)H1c2—C2—C3105.38
O1—K1—O367.58 (3)H1c2—C2—C4110.8
O1i—K1—O2128.05 (4)C3—C2—C4110.4 (3)
O1i—K1—O2ii100.88 (4)H1c2a—C2a—C3a113.92
O1i—K1—O2iii130.31 (4)H1c2a—C2a—C4a107.85
O1i—K1—O3iv52.67 (3)C3a—C2a—C4a109.5 (12)
O1i—K1—O373.84 (3)C2—C3—H1c3109.47
O2—K1—O2ii89.24 (4)C2—C3—H2c3109.47
O2—K1—O2iii89.30 (4)C2—C3—H3c3109.47
O2—K1—O3iv84.87 (2)H1c3—C3—H2c3109.47
O2—K1—O383.88 (2)H1c3—C3—H3c3109.47
O2ii—K1—O2iii112.43 (4)H2c3—C3—H3c3109.47
O2ii—K1—O3iv70.78 (3)C2a—C3a—H1c3a109.47
O2ii—K1—O3165.48 (4)C2a—C3a—H2c3a109.47
O2iii—K1—O3iv173.34 (3)C2a—C3a—H3c3a109.47
O2iii—K1—O380.35 (4)H1c3a—C3a—H2c3a109.47
O3iv—K1—O395.81 (3)H1c3a—C3a—H3c3a109.47
K1—O1—K1i87.96 (4)H2c3a—C3a—H3c3a109.47
K1—O2—K1ii90.76 (4)C2—C4—H1c4109.47
K1—O2—K1iii90.70 (4)C2—C4—H2c4109.47
K1ii—O2—K1iii112.43 (4)C2—C4—H3c4109.47
K1—O3—K1v95.806 (12)H1c4—C4—H2c4109.47
K1—O3—K1i93.77 (5)H1c4—C4—H3c4109.47
K1—O3—K1vi170.43 (4)H2c4—C4—H3c4109.47
K1v—O3—K1i170.43 (4)H1c2—C4a—C2a66.18
K1v—O3—K1vi74.62 (3)C2a—C4a—H1c4a109.47
K1i—O3—K1vi95.806 (12)C2a—C4a—H2c4a109.47
H1o3—O3—H1o3i105.0 (16)C2a—C4a—H3c4a109.47
O1—C1—O2124.40 (13)H1c4a—C4a—H2c4a109.47
O1—C1—C2119.45 (14)H1c4a—C4a—H3c4a109.47
O1—C1—C2a109.8 (4)H2c4a—C4a—H3c4a109.47
O2—C1—C2116.04 (14)
Symmetry codes: (i) x+1, y, z+1/2; (ii) x+3/2, y1/2, z+1/2; (iii) x+3/2, y+1/2, z+1/2; (iv) x, y1, z; (v) x, y+1, z; (vi) x+1, y+1, z+1/2.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O3—H1O3···O1v0.83 (2)1.92 (2)2.7358 (17)167 (2)
O3—H1O3i···O1vi0.83 (2)1.92 (2)2.7358 (17)167 (2)
Symmetry codes: (i) x+1, y, z+1/2; (v) x, y+1, z; (vi) x+1, y+1, z+1/2.
Overview of structural types observed in isobutyrates top
NumberRefcodeReference
Water-free clusters
1OHOXUFMalaestean et al. (2009)
2GEWFULMalaestean et al. (2013a)
3NAGQUICoker et al. (2004)
Clusters interconnected by water molecules
4NAGQUICoker et al. (2004)
Water-free columns
5PENJUNIlina et al. (1992)
6TAHXORBoyle et al. (2010)
7TAHXOR01Bierke & Meyer (2008)
8TAHXOR02Kotal et al. (2010)
Columns interconnected by water molecules
9MECVAUSkelton & Deacon (2017)
10XALZANMalaestean et al. (2012)
Water-free layered structures
11KELKOESkelton & Deacon (2017)
12LUHGOKYuranov & Dunaeva (1989)
Structures with layers interconnected by water molecules
13VIQTOGMalaestean et al. (2013b)
14POSCIJTroyanov et al. (1993)
15SAJMUOFischer et al. (2017)
Notes: 1: Bis(µ4-oxo)dodecakis(µ3-isobutyrato)hexakis(µ2-isobutyrato)bis(isobutyric acid)-bis(propanol)-octa-manganese(II)-di-manganese(III) dihydrate; 2: hexakis(µ3-isobutyrato)hexakis(µ2-isobutyrato)hexakis(2-methylpropanoic acid)hexamanganese; 3: hexakis[bis(µ2-2-methylpropanoato)(2-methylpropanoic acid)magnesium]; 4: bis(µ4-oxo)dodecakis(µ3-isobutyrato)hexakis(µ2-isobutyrato)bis(isobutyric acid)bis(propanol)octamanganese(II)dimanganese(III) propanol solvate; 5: bis[(µ3-isobutyrylato-O,O')(µ2-isobutyrylato)copper(II)]; 6: catena-[tetrakis(µ2-2-methylpropanoato)dizinc]; 7: catena-[tetrakis(µ2-2-methylpropanoato)dizinc); 8: catena-[tetrakis(µ2-2-methylpropanoato)dizinc); 9: catena-[tetrakis(µ-2-methylpropanoato)bis(2-methylpropanoato)triaquadilanthanum(III) hydrate; 10: catena-[tetrakis(µ2-isobutyrato-O,O,O')bis(isobutyrato-O,O')triaquadicerium ethanol solvate]; 11: catena-[hexakis(µ-methylpropanoato)dilutetium]; 12: diaquabis(isobutyrato)dioxouranium(VI); 13] catena-[(µ2-2-methylpropanoato)(2-methylpropanoato)triaquamagnesium monohydrate]; 14: catena-[hexakis(µ2-isobutyrato)aquadierbium monohydrate]; 15] catena-[bis(µ-2-methylpropanoato)(µ-aqua)cobalt(II) monohydrate].
Cmethylene—Cmethyl and Cmethyl—Cmethyl intermolecular distances (Å) in the title structure of up to 4.5 Å top
C2, C2a correspond to the methylene carbon atoms while C3, C3a and C4, C4a correspond to methyl atoms.
C2···C4v4.251 (6)C3···C4vi3.992 (6)
C3···C4viii4.136 (6)C4···C4ix3.984 (5)
C3a···C4avii4.17 (2)C3a···C4avi4.25 (2)
C4a···C4aix4.31 (2)C2a···C3ax3.97 (2)
C2a···C4ax3.884 (19)C2···C2av3.960 (10)
C2···C4ax4.479 (16)C3···C2av4.128 (11)
C3···C3ax4.44 (2)C3···C4avii4.000 (15)
C3···C4avi4.207 (15)C4···C3ax4.48 (2)
C4···C3axi4.087 (14)C4···C3aviii4.43 (3)
C4···C4ax3.767 (17)C4···C4aix4.051 (17)
Symmetry codes: (v) x, y + 1, z; (vi) x - 1/2, y + 1/2, z; (vii) x - 1/2, y - 1/2, z; (viii) -x + 1, -y, -z; (ix) -x + 3/2, -y + 1/2, -z; (x) x, y - 1, z; (xi) x + 1/2, y - 1/2, z; (xii) x + 1/2, y + 1/2, z.
 

Acknowledgements

Dr Ivana Císařová from the Faculty of Science is thanked for the generous measurement of the sample.

Funding information

The authors express their gratitude for the support provided by Project NPU I – LO1603 of the Ministry of Education of the Czech Republic to the Institute of Physics of the Academy of Sciences of the Czech Republic.

References

First citationBierke, T. & Meyer, G. (2008). Thesis, Universität zu Köln, Germany.  Google Scholar
First citationBoyle, T. J., Raymond, R., Boye, D. M., Ottley, L. A. M. & Lu, P. (2010). Dalton Trans. 39, 8050–8063.  CSD CrossRef CAS PubMed Google Scholar
First citationBrandenburg, K. (2005). DIAMOND. Crystal Impact GbR, Bonn, Germany.  Google Scholar
First citationBrese, N. E. & O'Keeffe, M. (1991). Acta Cryst. B47, 192–197.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationBruker (2017). Instrument Service, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationCoker, E. N., Boyle, T. J., Rodriguez, M. A. & Alam, T. M. (2004). Polyhedron, 23, 1739–1747.  Web of Science CSD CrossRef CAS Google Scholar
First citationDumbleton, J. H. & Lomer, T. R. (1965). Acta Cryst. 19, 301–307.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationFábry, J. & Samolová, E. (2020). Acta Cryst. E76, 1508–1513.  CSD CrossRef IUCr Journals Google Scholar
First citationFerloni, P., Sanesi, M. & Franzosini, P. (1975). Z. Naturforsch. Teil A, 30, 1447–1457.  CrossRef Web of Science Google Scholar
First citationFischer, A. I., Gurzhiy, V. V., Aleksandrova, J. V. & Pakina, M. I. (2017). Acta Cryst. E73, 318–321.  CSD CrossRef IUCr Journals Google Scholar
First citationGilli, G. & Gilli, P. (2009). The Nature of the Hydrogen Bond: Outline of a comprehensive hydrogen bond theory, p. 61. Oxford University Press.  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 citationHeyman, J. B., Chen, C.-H. & Foxman, B. M. (2020). Cryst. Growth Des. 20, 330–336.  Web of Science CSD CrossRef CAS Google Scholar
First citationIlina, E. G., Troyanov, S. I. & Dunaeva, K. M. (1992). Koord. Khim. 18, 882–890.  CAS Google Scholar
First citationKotal, A., Paira, T. K., Banerjee, S. & Mandal, T. K. (2010). Langmuir, 26, 6576–6582.  CSD CrossRef CAS PubMed Google Scholar
First citationMalaestean, I. L., Ellern, A., Baca, S. & Kögerler, P. (2012). Chem. Commun. 48, 1499–1501.  Web of Science CSD CrossRef CAS Google Scholar
First citationMalaestean, I. L., Ellern, A. & Kögerler, P. (2013a). Eur. J. Inorg. Chem., pp. 1635–1638.  Google Scholar
First citationMalaestean, I. L., Schmitz, S., Ellern, A. & Kögerler, P. (2013b). Acta Cryst. C69, 1144–1146.  CSD CrossRef IUCr Journals Google Scholar
First citationMalaestean, I. L., Speldrich, M., Ellern, A., Baca, S. G., Ward, M. & Kögerler, P. (2009). Eur. J. Inorg. Chem. pp. 4209–4212.  CSD CrossRef Google Scholar
First citationMartínez Casado, F. J., Ramos Riesco, M., da Silva, I., Labrador, A., Redondo, M. I., García Pérez, M. V., López-Andrés, S. & Rodríguez Cheda, J. A. (2010). J. Phys. Chem. B114, 10075–10085.  Google Scholar
First citationMartínez Casado, F. J., Riesco, M. R., García Pérez, M. V., Redondo, M. I., López-Andrés, S. & Rodríguez Cheda, J. A. (2009). J. Phys. Chem. B, 113, 12896–12902.  Web of Science PubMed Google Scholar
First citationMirnaya, T. A., Polishchuk, A. P., Molochaeva, V. I. & Tolochko, A. S. (1991). Kristallografiya, 36, 377–383.  CAS Google Scholar
First citationPetříček, V., Dušek, M. & Palatinus, L. (2014). Z. Kristallogr. 229, 345–352.  Google Scholar
First citationSheldrick, G. M. (2015). Acta Cryst. A71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSkelton, B. W. & Deacon, G. B. (2017). CSD Communication (refcodes KELKOE, MECVAU). CCDC, Cambridge, England.  Google Scholar
First citationStadnicka, K. & Glazer, A. M. (1980). Acta Cryst. B36, 2977–2985.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationTroyanov, S. I., Mitrofanova, N. D. & Gorbacheva, M. V. (1993). Koord. Khim. 19, 868–870.  CAS Google Scholar
First citationYuranov, I. A. & Dunaeva, K. M. (1989). Koord. Khim. 15, 845–847.  CAS Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

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