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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

μ-Methyl­ene-bis­­[di­bromido(di­ethyl ether-κO)aluminium(III)]: crystal structure and chemical exchange in solution

crossmark logo

aDepartment of Chemistry, Howard University, 525 College Street NW, Washington DC 20059, USA, and bChemistry Division, Code 6123, Naval Research Laboratory, 4555 Overlook Av, SW, Washington DC 20375-5342, USA
*Correspondence e-mail: rbutcher99@yahoo.com

Edited by V. Jancik, Universidad Nacional Autónoma de México, México (Received 11 December 2020; accepted 19 May 2021; online 21 May 2021)

In the title compound, [Al2Br4(CH2)(C4H10O)2], the mol­ecule lies on a crystallographic twofold axis passing through the bridging C atom. Each AlIII atom is four-coordinate, being bonded to two bromide ions, bridging the CH2 group as well as the oxygen atom of a diethyl ether ligand in a slightly distorted tetra­hedral arrangement with angles ranging from 101.52 (8) to 116.44 (5)°. The Al—CH2—Al angle, 118.4 (2)°, is the smallest observed for a structure where this moiety is not part of a ring. In the crystal, weak C—H⋯Br inter­actions, characterized as R22(12) rings, link the mol­ecules into ribbons in the [101] direction. The title compound is monomeric and coordinatively saturated in the solid state, as each aluminum is four-coordinate, but in solution the ether mol­ecules from either or both Al atoms can dissociate, and would be expected to rapidly exchange, and this is supported by NMR data.

1. Chemical context

There is great current inter­est in the chemistry of reduced aluminum (Klemp et al., 2001[Klemp, C., Bruns, M., Gauss, J., Häussermann, U., Stösser, G., van Wüllen, L., Jansen, M. & Schnöckel, H. (2001). J. Am. Chem. Soc. 123, 9099-9106.], Bonyhady et al., 2018[Bonyhady, S. J., Collis, D., Holzmann, N., Edwards, A. J., Piltz, R. O., Frenking, G., Stasch, A. & Jones, C. (2018). Nat. Commun. 9, 3079.]) and aluminum carbon (carbaalanes) clusters (Stasch et al., 2002[Stasch, A., Ferbinteanu, M., Prust, J., Zheng, W., Cimpoesu, F., Roesky, H. W., Magull, J., Schmidt, H.-G. & Noltemeyer, M. (2002). J. Am. Chem. Soc. 124, 5441-5448.]; Uhl & Roesky, 2002[Uhl, W. & Roesky, H. W. (2002). Inorganic Chemistry Highlights, pp. 357-389.]; Kumar et al., 2004[Kumar, S. S., Rong, J., Singh, S., Roesky, H. W., Vidovic, D., Magull, J., Neculai, D., Chandrasekhar, V. & Baldus, M. (2004). Organo­met­allics, 23, 3496-3500.]) as well as aluminum–carbon nanoparticles (Diaz-Droguett et al., 2020[Diaz-Droguett, D. E., Ramos-Moore, E., Roble, M. & Mücklich, F. (2020). Ceram. Int. 46, 20456-20462.]) because of their inter­esting structural chemistry and many theoretical studies have been carried out on potential derivatives and as analogs of the better known boron examples (Attia et al., 2017[Attia, A. A. A., Lupan, A. & King, R. B. (2017). Organometallics, 36, 1019-1026.]). This has lead to a renaissance in the chemistry of aluminum (Roesky, 2004[Roesky, H. W. (2004). Inorg. Chem. 43, 7284-7293.]). In view of this chemistry, there is a need for easily prepared precursors for the synthesis of these reduced aluminum and carbaalane clusters, and this is the motivation behind preparing organometallics with two or more Al atoms on a carbon atom. The synthesis of methyl­ene bis­(aluminum halides) has been described before (Ort & Mottus, 1973[Ort, M. R. & Mottus, E. H. (1973). J. Organomet. Chem. 50, 47-52.]; Lehmkuhl & Schäfer, 1966[Lehmkuhl, H. & Schäfer, R. (1966). Tetrahedron Lett. 7, 2315-2320.]).

[Scheme 1]

2. Structural commentary

In the structure of the title compound, [Al2(C9H22Br4O2)] (1), the mol­ecule lies on a crystallographic twofold axis passing through C1 (see Fig. 1[link]). Each Al atom is four-coordinate, being bonded to two bromide ions and the bridging CH2 group as well as the oxygen of a diethyl ether ligand in a slightly distorted tetra­hedral arrangement (τ4 = 0.907; Okuniewski et al., 2015[Okuniewski, A., Rosiak, D., Chojnacki, J. & Becker, B. (2015). Polyhedron, 90, 47-57.]) with angles ranging from 101.52 (8) to 116.44 (5)° (see Table 1[link]). In the literature there are eight structures containing an AlBr2 fragment coordinated to a diethyl ether ligand (LOCMEY, Yanagisawa et al., 2018[Yanagisawa, T., Mizuhata, Y. & Tokitoh, N. (2018). Heteroat. Chem. 29, e21465.]; NOJYIW, Lips et al., 2014[Lips, F., Fettinger, J. C. & Power, P. P. (2014). Polyhedron, 79, 207-212.]; QQQGXV, QQQGYA, Semenenko et al., 1973[Semenenko, K. N., Lobkovskii, E. B. & Fokin, V. N. (1973). Zh. Neorg. Khim. 18, 2718-2722.]; RABCOM, Wehmschulte et al., 1996[Wehmschulte, R. J., Grigsby, W. J., Schiemenz, B., Bartlett, R. A. & Power, P. P. (1996). Inorg. Chem. 35, 6694-6702.]; TEXNIV, Agou et al., 2012[Agou, T., Nagata, K., Sakai, H., Furukawa, Y. & Tokitoh, N. (2012). Organometallics, 31, 3806-3809.]; YANKON, Petrie et al., 1993[Petrie, M. A., Power, P. P., Dias, H. V. R., Ruhlandt-Senge, K., Waggoner, K. M. & Wehmschulte, R. J. (1993). Organometallics, 12, 1086-1093.]; YERLUD, Quillian et al., 2006[Quillian, B., Wang, Y., Wei, P., Handy, A. & Robinson, G. H. (2006). J. Organomet. Chem. 691, 3765-3770.]). In each of these structures, there is both a longer and shorter Al—Br bond distance [average Al—Br distances of 2.315 (18) and 2.30 (2) Å] with an average Al—O distance of 1.874 (14) Å. The comparable distances in 1 are 2.3046 (10), 2.3029 (9) and 1.881 (2) Å.

Table 1
Selected geometric parameters (Å, °)

Al—O1 1.881 (2) Al—Br2 2.3029 (9)
Al—C1 1.927 (2) Al—Br1 2.3046 (10)
       
O1—Al—C1 110.42 (13) O1—Al—Br1 101.52 (8)
O1—Al—Br2 101.60 (7) C1—Al—Br1 116.44 (5)
C1—Al—Br2 114.90 (9) Br2—Al—Br1 110.07 (4)
[Figure 1]
Figure 1
Mol­ecular diagram showing the atom labeling (symmetry operation to generate the complete mol­ecule, −x, y, [{1\over 2}] − z). Atomic displacement parameters are at the 30% level.

As indicated below, there are many instances of structures containing the Al–CH2–Al fragment but only one which combines this fragment along with aluminum–halogen bonding (Uhl & Layh, 1991[Uhl, W. & Layh, M. (1991). J. Organomet. Chem. 415, 181-190.]). In this structure {[(Me3Si)2CHAlCl]2CH2}2, this moiety is not isolated but part of a ring in an adamantanoid cage, which would influence both its bond lengths and angles. However, there are ten instances (BELLAH, BELLEL, BELLIP, BELLOW, BELLUP (Uhl et al., 2012a[Uhl, W., Rösener, C., Layh, M. & Hepp, A. (2012a). Z. Anorg. Allg. Chem. 638, 1746-1754.]); JEZFID (Layh & Uhl, 1990[Layh, M. & Uhl, W. (1990). Polyhedron, 9, 277-282.]); JUWMOD (Uhl et al., 1993[Uhl, W., Koch, M. & Vester, A. (1993). Z. Anorg. Allg. Chem. 619, 359-366.]); PENSEI (Uhl et al., 2012b[Uhl, W., Rösener, C., Stefaniak, C., Choy, T., Jasper-Peter, B., Kösters, J., Layh, M. & Hepp, A. (2012b). Z. Naturforsch. B: J. Chem. Sci. 67, 1081-1090.]); WOZJUQ, WOZKAX (Knabel et al., 2002[Knabel, K., Nöth, H. & Seifert, T. (2002). Z. Naturforsch. Teil B, 57, 830-834.]) where the metrical parameters of the Al–CH2–Al fragment are not influenced by being part of a ring. In these structures, apart from JEZFID (Layh & Uhl, 1990[Layh, M. & Uhl, W. (1990). Polyhedron, 9, 277-282.]) and WOZJUQ (Knabel et al., 2002[Knabel, K., Nöth, H. & Seifert, T. (2002). Z. Naturforsch. Teil B, 57, 830-834.]), there are two independent Al—C bond lengths, which average 2.003 and 1.922 Å, with an overall average Al—C—Al bond angle of 132.5°. As a result of the unconstrained nature of this angle, it varies over a wide range from 126.3 to 144.4° and the value depends on the steric bulk of the Al substituents. In the smallest value in the list [BELLOV, 126.29 (13)°; Uhl et al., 2012a[Uhl, W., Rösener, C., Layh, M. & Hepp, A. (2012a). Z. Anorg. Allg. Chem. 638, 1746-1754.]], the substituents attached to Al are (tri­methyl­sil­yl)meth­yl moieties, while the largest [JUWMOD, 144.4 (2)°; Uhl et al., 1993[Uhl, W., Koch, M. & Vester, A. (1993). Z. Anorg. Allg. Chem. 619, 359-366.]] has a neopentyl as well as two [bis­(tri­methyl­sil­yl)meth­yl] groups attached. There are two structures, QQQGXV and QQQGYA (Semenenko et al., 1973[Semenenko, K. N., Lobkovskii, E. B. & Fokin, V. N. (1973). Zh. Neorg. Khim. 18, 2718-2722.]), which only have Br3 and Br2H as substituents on the Al, but the angles cannot be calculated since the coordinates are not available. In 1, which lacks this steric bulk and where atom C1 lies on a crystallographic twofold axis, these values are 1.927 (2) Å and 118.4 (2)°, respectively. This latter value reflects this lack of steric bulk in the groups attached to the Al atoms.

3. Supra­molecular features

As shown in Fig. 2[link], there are weak C—H⋯Br inter­actions, which link the mol­ecules into ribbons in the [101] direction (see Table 2[link]). In graph-set notation (Etter et al., 1990[Etter, M. C., MacDonald, J. C. & Bernstein, J. (1990). Acta Cryst. B46, 256-262.]), these inter­actions can be characterized as R22(12) rings and this is shown in Fig. 3[link]. These inter­actions can be highlighted in a Hirshfeld fingerprint plot as shown in Fig. 4[link] (Spackman & Jayatilaka, 2009[Spackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19-32.]), which shows these features. If this is expanded to take inter­actions beyond the van der Waals radii sum cutoff (Desiraju & Steiner, 1999[Desiraju, G. R. & Steiner, T. (1999). Editors. The Weak Hydrogen Bond in Structural Chemistry and Biology. New York: Oxford University Press Inc.]; Desiraju, 2011a[Desiraju, G. R. (2011a). Angew. Chem. Int. Ed. 50, 52-59.],b[Desiraju, G. R. (2011b). Cryst. Growth Des. 11, 896-898.]), this plot indicates that these weak C—H⋯Br inter­actions dominate the packing and make up 52.6% of all inter­molecular inter­actions.

Table 2
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
C2—H2A⋯Br1 0.99 2.98 3.481 (3) 113
C5—H5B⋯Br1i 0.98 3.10 3.871 (4) 136
Symmetry code: (i) [-x+{\script{1\over 2}}, -y+{\script{3\over 2}}, -z+1].
[Figure 2]
Figure 2
Packing diagram of 1 viewed from the [010] direction.
[Figure 3]
Figure 3
Diagram showing the C—H⋯Br inter­actions (as dashed lines) that link the mol­ecules into ribbons via the formation of R22(12) rings (symmetry operation, [{1\over 2}] − x, [{3\over 2}] − y, 1 − z).
[Figure 4]
Figure 4
Hirshfeld surface plot highlighting the C—H⋯Br inter­actions, which make up 52.6% of all inter­actions.

4. Database survey

A search of the Cambridge Structural Database [CSD version 5.41 (November 2019); Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]] for fragments based on the structure of 1 revealed there are eight structures in the literature containing an AlBr2 fragment coordinated to a diethyl ether ligand (LOCMEY, Yanagisawa et al., 2018[Yanagisawa, T., Mizuhata, Y. & Tokitoh, N. (2018). Heteroat. Chem. 29, e21465.]; NOJYIW, Lips et al., 2014[Lips, F., Fettinger, J. C. & Power, P. P. (2014). Polyhedron, 79, 207-212.]; QQQGXV, QQQGYA, Semenenko et al., 1973[Semenenko, K. N., Lobkovskii, E. B. & Fokin, V. N. (1973). Zh. Neorg. Khim. 18, 2718-2722.]; RABCOM, Wehmschulte et al., 1996[Wehmschulte, R. J., Grigsby, W. J., Schiemenz, B., Bartlett, R. A. & Power, P. P. (1996). Inorg. Chem. 35, 6694-6702.]; TEXNIV, Agou et al., 2012[Agou, T., Nagata, K., Sakai, H., Furukawa, Y. & Tokitoh, N. (2012). Organometallics, 31, 3806-3809.]; YANKON, Petrie et al., 1993[Petrie, M. A., Power, P. P., Dias, H. V. R., Ruhlandt-Senge, K., Waggoner, K. M. & Wehmschulte, R. J. (1993). Organometallics, 12, 1086-1093.]; YERLUD, Quillian et al., 2006[Quillian, B., Wang, Y., Wei, P., Handy, A. & Robinson, G. H. (2006). J. Organomet. Chem. 691, 3765-3770.]). There were 99 examples containing the Al–CH2–Al fragment, of which there are ten instances (BELLAH, BELLEL, BELLIP, BELLOW, BELLUP (Uhl et al., 2012a[Uhl, W., Rösener, C., Layh, M. & Hepp, A. (2012a). Z. Anorg. Allg. Chem. 638, 1746-1754.]); JEZFID (Layh & Uhl, 1990[Layh, M. & Uhl, W. (1990). Polyhedron, 9, 277-282.]); JUWMOD (Uhl et al., 1993[Uhl, W., Koch, M. & Vester, A. (1993). Z. Anorg. Allg. Chem. 619, 359-366.]); PENSEI (Uhl et al., 2012b[Uhl, W., Rösener, C., Stefaniak, C., Choy, T., Jasper-Peter, B., Kösters, J., Layh, M. & Hepp, A. (2012b). Z. Naturforsch. B: J. Chem. Sci. 67, 1081-1090.]); WOZJUQ, WOZKAX (Knabel et al., 2002[Knabel, K., Nöth, H. & Seifert, T. (2002). Z. Naturforsch. Teil B, 57, 830-834.]) where the metrical parameters of the Al–CH2–Al fragment are not influenced by being part of a ring.

5. Synthesis and crystallization

Aluminum wire, cut into small pieces (3.19 g), was added slowly over several days to a stirred, dry CH2Br2 (50 mL) under N2 by inserting the wire through a hole in a rubber septum. After the aluminum had reacted, the mixture was filtered inside an N2 flow dry box and the solids were collected and pumped dry. A total of 20.22 g (96% based on Al) was isolated. A portion of this white solid was dissolved in Et2O and allowed to slowly evaporate inside the dry box to produce crystals of the title compound. IR (neat smeared on KBr plates, cm−1): [3002.90, 2982.63, 2871.84, 2964.61, 2935.43, 2920.37, 2850.50] (m, C—H str), 2213.10 (w), 1635.50 (w), 1463.47 (m), 1442.72 (m), 1390.14 (s), 1326.09 (m), 1281.45 (w), 1260.61 (m), 1189.28 (m), 1146.67 (m), 1088.57 (m), 999.64 (s), 985.03 (s), 904.06 (w), 879.11 (s), 827.67 (m), 796.06 (w), 763.57 (s), 723.29 (m), 606.38 (s), 545.18 (s), 530.16 (s), 463.05 (w). The NMR solvents were dried from sodium–potassium alloy. NMR spectra were recorded in C6D6 solution in flame-sealed tubes and were found to be concentration dependent. Proton spectra were recorded at 400 MHz on three different concentrations, and two samples of the inter­mediate concentration with added ether, and are displayed in Fig. 5[link]. 13C spectrum (C6D6, 100 MHz): δ 1.34 (CH2, sharp), −1.46 ppm (CH2, broad, HHLW ≃ 150 Hz). 27Al spectrum (C6D6, 104 MHz): δ 93 (sharp), 132 ppm (broad, HHLW ≃ 4000 Hz).

[Figure 5]
Figure 5
1H NMR spectra of the title compound in C6D6 at three different concentrations (bottom three spectra), and at an inter­mediate concentration with added ether (top two spectra). The CH2 group attached to Al has peaks A, B, C, and D, which are concentration dependent, and an expanded view from δ 1 to −1 ppm is shown in the lower part of the figure. The concentration of the small peak at 2.5 ppm (probably OH) is invariant in all samples and is undoubtedly due to hydrolysis caused by the release of a small amount of water during flame sealing of the NMR tubes.

Safety Note:

This reaction should be carried out with caution as when finely divided Al flakes were used instead of Al wire, an explosion occurred.

6. Chemical exchange in solution

The title compound (1) is monomeric and coordinatively saturated in the solid state, as each aluminum is four-coord­inate, but in solution, the ether mol­ecules from either or both Al atoms can dissociate and would be expected to exchange rapidly. Once an ether mol­ecule dissociates, the aluminum atom can regain four-coordination by association to a bromine atom from the other half of the same or another mol­ecule. In the C6D6 solution, there are four main proton NMR peaks visible for the CH2 moieties on aluminum, as shown in Fig. 5[link], and those peaks are labeled A, B, C, and D. Both the relative amounts and chemical shifts of peaks AD are concentration dependent. Additionally, the NMR peaks for the ether moieties are dependent on concentration as well, as is most obvious at the lowest concentration (where the ether CH2 peak splits), and adding additional ether to the solution does affect the spectra, as also shown in Fig. 5[link]. The unsolvated parent compound, CH2(AlBr2)2, has extremely low solubility in non-coordinating solvents such as C6D6, as one would expect if it is polymeric. While the structures of unsolvated compounds of this type are unknown, association through Al2Br2 rings is common, and this compound can easily form such rings on each end linking into an extended structure.

One can determine some information as to the identity of the peaks from the concentration dependence of the spectra. If there is an exchange process between different degrees of association (for example between monomer and trimer), the ratio of oligomerization between the species can be determined by the slope of a ln–ln plot of the molar concentrations represented by each NMR peak (Purdy et al., 1987[Purdy, A. P, Wells, R. L., McPhail, A. T. & Pitt, C. G. (1987) Organometallics, 6, 2099-2105.]). Fig. 6[link] shows a natural ln–ln plot for the integral fraction of the CH2 NMR peaks multiplied by the absolute concentration of the title compound in solution, for all six binary combinations of peaks AD, with the linear equation between the points displayed on the chart. All combinations involving only peaks A, C, and D have R2 factors near 1, showing a high linear correlation. The species with the NMR peak C clearly has three times the degree of association of A, and D has 2.5 times the degree of association as A. However, all combinations involving peak B with A, C, or D do not have as good a linear correlation, but do show that and A and B have approximately the same degree of association. Fig. 7[link] displays a ln–ln plot for the concentrations of peaks AD against the ether concentration for three solutions of approximately the same concentration of the title compound with varying amounts of ether. Clearly, peak B correlates positively to the ether concentration, and the relative amounts of peaks A, C, and D have a slightly negative correlation to the concentration of ether. Therefore, we conclude that B is for a solution species that is more coordinatively saturated by ether than A, C, or D, and is probably the title compound. A is probably formed by the dissociation of a single ether mol­ecule, C is its trimer, and D is a partially associated trimer. Fig. 8[link] illustrates some possible structures of monomeric and trimeric species, with varying degrees of ether solvation, although the drawings of trimers do not exhaust the possible structures that may exist. The NMR data do not allow definite structural conclusions to be drawn for the trimers. While substantial precedent exists for compounds associated through Al2Br2 rings, and an example exists for a four-membered Al–CH2–Al–Br ring (PENSOS; Uhl et al. 2012a[Uhl, W., Rösener, C., Layh, M. & Hepp, A. (2012a). Z. Anorg. Allg. Chem. 638, 1746-1754.]), six- and eight-membered aluminum–halogen (AlX)n rings are mostly known for X=F, although a structurally constrained Cl example does exist (GOTNEI; Tschinkl et al. 1999[Tschinkl, M., Gabbaï, F. P. & Bachman, R. E. (1999). Chem. Commun. pp. 1367-1368.]). An example exists for a linearly associated –CH2–AlBr3–AlBr3 moiety (KIXBEA; Ménard et al. 2013[Ménard, G., Tran, L., McCahill, J. S. J., Lough, A. J. & Stephan, D. W. (2013). Organometallics, 32, 6759-6763.]), which opens the possibility that a partially associated trimer could have a single dative bond in place of Al2Br2 or Al3Br3 rings.

[Figure 6]
Figure 6
ln–ln plots of the concentrations of the mol­ecules represented by the CH2—Al peaks in the proton spectra. The concentration is calculated from the integral fraction of those CH2 resonances multiplied by the total concentration of CH2(AlBr2OEt2)2 dissolved. The slope of the line is the ratio of the degree of association of the species in solution.
[Figure 7]
Figure 7
ln–ln plots of the total ether concentration on the x-axis and the concentration of the species represented by the CH2—Al peaks on the y-axis for the three samples with approximately equal total concentration of CH2(AlBr2OEt2)2. The relative amount of the species with peak B increases with ether concentration, while the other peaks decrease.
[Figure 8]
Figure 8
Drawings of some possible structures of monomeric and trimeric aggregates of 1, with varying degrees of ether coordination. The possibilities exhibited here are not exhaustive.

7. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 3[link]. For the CH2 bridging group, the H-atom position was refined isotropically while the other H atoms were refined in idealized positions using a riding model with atomic displacement parameters of Uiso(H) = 1.2Ueq(C) [1.5Ueq(C) for CH3], with C—H distances ranging from 0.98 to 0.99 Å.

Table 3
Experimental details

Crystal data
Chemical formula [Al2Br4(CH2)(C4H10O)2]
Mr 535.86
Crystal system, space group Monoclinic, C2/c
Temperature (K) 100
a, b, c (Å) 8.3872 (6), 12.1039 (6), 18.1504 (12)
β (°) 95.646 (3)
V3) 1833.7 (2)
Z 4
Radiation type Mo Kα
μ (mm−1) 8.87
Crystal size (mm) 0.31 × 0.25 × 0.08
 
Data collection
Diffractometer Bruker APEXII CCD
Absorption correction Multi-scan (SADABS; Sheldrick, 1996[Sheldrick, G. M. (1996). SADABS. University of Göttingen, Germany.])
Tmin, Tmax 0.291, 0.747
No. of measured, independent and observed [I > 2σ(I)] reflections 10600, 2027, 1736
Rint 0.073
(sin θ/λ)max−1) 0.641
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.031, 0.076, 1.07
No. of reflections 2027
No. of parameters 83
H-atom treatment H atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3) 0.74, −0.73
Computer programs: APEX2 (Bruker, 2005[Bruker (2005). APEX2. Bruker AXS Inc., Madison, Wisconsin, USA.]), SAINT (Bruker, 2002[Bruker (2002). SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXT (Sheldrick 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2018/3 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), and SHELXTL (Sheldrick 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]).

Supporting information


Computing details top

Data collection: APEX2 (Bruker, 2005); cell refinement: SAINT (Bruker, 2002); data reduction: SAINT (Bruker, 2002); program(s) used to solve structure: SHELXT (Sheldrick 2015a); program(s) used to refine structure: SHELXL2018/3 (Sheldrick, 2015b); molecular graphics: SHELXTL (Sheldrick 2008); software used to prepare material for publication: SHELXTL (Sheldrick 2008).

µ-Methylene-bis[dibromido(diethyl ether-κO)aluminium(III)] top
Crystal data top
[Al2Br4(CH2)(C4H10O)2]F(000) = 1032
Mr = 535.86Dx = 1.941 Mg m3
Monoclinic, C2/cMo Kα radiation, λ = 0.71073 Å
a = 8.3872 (6) ÅCell parameters from 4301 reflections
b = 12.1039 (6) Åθ = 3.1–32.5°
c = 18.1504 (12) ŵ = 8.87 mm1
β = 95.646 (3)°T = 100 K
V = 1833.7 (2) Å3Plate, colorless
Z = 40.31 × 0.25 × 0.08 mm
Data collection top
Bruker APEXII CCD
diffractometer
1736 reflections with I > 2σ(I)
φ and ω scansRint = 0.073
Absorption correction: multi-scan
(SADABS; Sheldrick, 1996)
θmax = 27.1°, θmin = 3.0°
Tmin = 0.291, Tmax = 0.747h = 910
10600 measured reflectionsk = 1515
2027 independent reflectionsl = 2323
Refinement top
Refinement on F2Primary atom site location: structure-invariant direct methods
Least-squares matrix: fullSecondary atom site location: difference Fourier map
R[F2 > 2σ(F2)] = 0.031Hydrogen site location: mixed
wR(F2) = 0.076H atoms treated by a mixture of independent and constrained refinement
S = 1.07 w = 1/[σ2(Fo2) + (0.0337P)2 + 0.0465P]
where P = (Fo2 + 2Fc2)/3
2027 reflections(Δ/σ)max = 0.001
83 parametersΔρmax = 0.74 e Å3
0 restraintsΔρmin = 0.73 e Å3
Special details top

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

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Al0.11619 (11)0.71423 (7)0.32918 (6)0.0103 (2)
Br10.02876 (4)0.77240 (3)0.42346 (2)0.01923 (12)
Br20.27035 (4)0.85713 (3)0.29158 (2)0.02154 (12)
O10.2720 (3)0.62389 (16)0.38074 (13)0.0112 (4)
C10.0000000.6327 (3)0.2500000.0131 (9)
H10.066 (4)0.589 (3)0.272 (2)0.016*
C20.2436 (4)0.5550 (2)0.44484 (19)0.0150 (7)
H2A0.2042550.6019470.4838720.018*
H2B0.3457540.5207760.4652740.018*
C30.1224 (4)0.4655 (3)0.4234 (2)0.0202 (8)
H3A0.1116730.4174890.4661160.030*
H3B0.1582430.4216690.3827480.030*
H3C0.0185010.4992970.4074550.030*
C40.4055 (4)0.5842 (3)0.34074 (19)0.0151 (7)
H4A0.3889250.6083720.2884250.018*
H4B0.4081610.5024820.3416000.018*
C50.5618 (4)0.6288 (3)0.3758 (2)0.0201 (7)
H5A0.6500790.5968340.3512750.030*
H5B0.5744900.6092710.4284430.030*
H5C0.5630120.7093610.3705530.030*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Al0.0122 (5)0.0062 (4)0.0126 (5)0.0002 (3)0.0013 (4)0.0007 (3)
Br10.0223 (2)0.01703 (18)0.0192 (2)0.00542 (12)0.00634 (15)0.00335 (13)
Br20.0235 (2)0.01272 (17)0.0279 (2)0.00778 (12)0.00023 (15)0.00757 (13)
O10.0110 (11)0.0102 (10)0.0127 (12)0.0012 (8)0.0025 (9)0.0028 (8)
C10.016 (2)0.0088 (19)0.014 (2)0.0000.0011 (19)0.000
C20.0167 (17)0.0151 (15)0.0128 (17)0.0009 (12)0.0001 (13)0.0045 (13)
C30.0199 (18)0.0159 (16)0.025 (2)0.0023 (13)0.0051 (15)0.0042 (14)
C40.0163 (17)0.0143 (14)0.0157 (17)0.0016 (12)0.0064 (14)0.0030 (13)
C50.0147 (18)0.0257 (17)0.0206 (19)0.0015 (13)0.0059 (14)0.0021 (14)
Geometric parameters (Å, º) top
Al—O11.881 (2)C2—H2B0.9900
Al—C11.927 (2)C3—H3A0.9800
Al—Br22.3029 (9)C3—H3B0.9800
Al—Br12.3046 (10)C3—H3C0.9800
O1—C21.470 (4)C4—C51.500 (5)
O1—C41.473 (4)C4—H4A0.9900
C1—H10.89 (4)C4—H4B0.9900
C1—H1i0.89 (4)C5—H5A0.9800
C2—C31.510 (4)C5—H5B0.9800
C2—H2A0.9900C5—H5C0.9800
O1—Al—C1110.42 (13)H2A—C2—H2B108.0
O1—Al—Br2101.60 (7)C2—C3—H3A109.5
C1—Al—Br2114.90 (9)C2—C3—H3B109.5
O1—Al—Br1101.52 (8)H3A—C3—H3B109.5
C1—Al—Br1116.44 (5)C2—C3—H3C109.5
Br2—Al—Br1110.07 (4)H3A—C3—H3C109.5
C2—O1—C4113.3 (2)H3B—C3—H3C109.5
C2—O1—Al124.38 (19)O1—C4—C5110.4 (3)
C4—O1—Al117.9 (2)O1—C4—H4A109.6
Al—C1—Ali118.4 (2)C5—C4—H4A109.6
Al—C1—H1105 (2)O1—C4—H4B109.6
Ali—C1—H1111 (2)C5—C4—H4B109.6
Al—C1—H1i111 (2)H4A—C4—H4B108.1
Ali—C1—H1i105 (2)C4—C5—H5A109.5
H1—C1—H1i107 (5)C4—C5—H5B109.5
O1—C2—C3111.1 (3)H5A—C5—H5B109.5
O1—C2—H2A109.4C4—C5—H5C109.5
C3—C2—H2A109.4H5A—C5—H5C109.5
O1—C2—H2B109.4H5B—C5—H5C109.5
C3—C2—H2B109.4
C1—Al—O1—C289.5 (2)Br1—Al—O1—C4170.52 (18)
Br2—Al—O1—C2148.1 (2)C4—O1—C2—C391.5 (3)
Br1—Al—O1—C234.6 (2)Al—O1—C2—C364.4 (3)
C1—Al—O1—C465.4 (2)C2—O1—C4—C585.6 (3)
Br2—Al—O1—C457.0 (2)Al—O1—C4—C5116.9 (3)
Symmetry code: (i) x, y, z+1/2.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
C2—H2A···Br10.992.983.481 (3)113
C5—H5B···Br1ii0.983.103.871 (4)136
Symmetry code: (ii) x+1/2, y+3/2, z+1.
 

Acknowledgements

RJB wishes to acknowledge the ONR Summer Faculty Research Program for funding in 2019 and 2020.

Funding information

Funding for this research was provided by: The Office of Naval Research.

References

First citationAgou, T., Nagata, K., Sakai, H., Furukawa, Y. & Tokitoh, N. (2012). Organometallics, 31, 3806–3809.  CSD CrossRef CAS Google Scholar
First citationAttia, A. A. A., Lupan, A. & King, R. B. (2017). Organometallics, 36, 1019–1026.  CrossRef CAS Google Scholar
First citationBonyhady, S. J., Collis, D., Holzmann, N., Edwards, A. J., Piltz, R. O., Frenking, G., Stasch, A. & Jones, C. (2018). Nat. Commun. 9, 3079.  CSD CrossRef PubMed Google Scholar
First citationBruker (2002). SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationBruker (2005). APEX2. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationDesiraju, G. R. (2011a). Angew. Chem. Int. Ed. 50, 52–59.  CrossRef CAS Google Scholar
First citationDesiraju, G. R. (2011b). Cryst. Growth Des. 11, 896–898.  CrossRef CAS Google Scholar
First citationDesiraju, G. R. & Steiner, T. (1999). Editors. The Weak Hydrogen Bond in Structural Chemistry and Biology. New York: Oxford University Press Inc.  Google Scholar
First citationDiaz-Droguett, D. E., Ramos-Moore, E., Roble, M. & Mücklich, F. (2020). Ceram. Int. 46, 20456–20462.  CAS Google Scholar
First citationEtter, M. C., MacDonald, J. C. & Bernstein, J. (1990). Acta Cryst. B46, 256–262.  CrossRef ICSD CAS Web of Science IUCr Journals 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 citationKlemp, C., Bruns, M., Gauss, J., Häussermann, U., Stösser, G., van Wüllen, L., Jansen, M. & Schnöckel, H. (2001). J. Am. Chem. Soc. 123, 9099–9106.  CSD CrossRef PubMed CAS Google Scholar
First citationKnabel, K., Nöth, H. & Seifert, T. (2002). Z. Naturforsch. Teil B, 57, 830–834.  CrossRef CAS Google Scholar
First citationKumar, S. S., Rong, J., Singh, S., Roesky, H. W., Vidovic, D., Magull, J., Neculai, D., Chandrasekhar, V. & Baldus, M. (2004). Organo­met­allics, 23, 3496–3500.  CSD CrossRef CAS Google Scholar
First citationLayh, M. & Uhl, W. (1990). Polyhedron, 9, 277–282.  CSD CrossRef CAS Google Scholar
First citationLehmkuhl, H. & Schäfer, R. (1966). Tetrahedron Lett. 7, 2315–2320.  CrossRef Google Scholar
First citationLips, F., Fettinger, J. C. & Power, P. P. (2014). Polyhedron, 79, 207–212.  CSD CrossRef CAS Google Scholar
First citationMénard, G., Tran, L., McCahill, J. S. J., Lough, A. J. & Stephan, D. W. (2013). Organometallics, 32, 6759–6763.  Google Scholar
First citationOkuniewski, A., Rosiak, D., Chojnacki, J. & Becker, B. (2015). Polyhedron, 90, 47–57.  Web of Science CSD CrossRef CAS Google Scholar
First citationOrt, M. R. & Mottus, E. H. (1973). J. Organomet. Chem. 50, 47–52.  CrossRef CAS Google Scholar
First citationPetrie, M. A., Power, P. P., Dias, H. V. R., Ruhlandt-Senge, K., Waggoner, K. M. & Wehmschulte, R. J. (1993). Organometallics, 12, 1086–1093.  CSD CrossRef CAS Web of Science Google Scholar
First citationPurdy, A. P, Wells, R. L., McPhail, A. T. & Pitt, C. G. (1987) Organometallics, 6, 2099–2105.  CSD CrossRef CAS Google Scholar
First citationQuillian, B., Wang, Y., Wei, P., Handy, A. & Robinson, G. H. (2006). J. Organomet. Chem. 691, 3765–3770.  CSD CrossRef CAS Google Scholar
First citationRoesky, H. W. (2004). Inorg. Chem. 43, 7284–7293.  CrossRef PubMed CAS Google Scholar
First citationSemenenko, K. N., Lobkovskii, E. B. & Fokin, V. N. (1973). Zh. Neorg. Khim. 18, 2718–2722.  CAS Google Scholar
First citationSheldrick, G. M. (1996). SADABS. University of Göttingen, Germany.  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. (2015a). Acta Cryst. A71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015b). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSpackman, M. A. & Jayatilaka, D. (2009). CrystEngComm, 11, 19–32.  Web of Science CrossRef CAS Google Scholar
First citationStasch, A., Ferbinteanu, M., Prust, J., Zheng, W., Cimpoesu, F., Roesky, H. W., Magull, J., Schmidt, H.-G. & Noltemeyer, M. (2002). J. Am. Chem. Soc. 124, 5441–5448.  CSD CrossRef PubMed CAS Google Scholar
First citationTschinkl, M., Gabbaï, F. P. & Bachman, R. E. (1999). Chem. Commun. pp. 1367–1368.  CSD CrossRef Google Scholar
First citationUhl, W., Koch, M. & Vester, A. (1993). Z. Anorg. Allg. Chem. 619, 359–366.  CSD CrossRef CAS Google Scholar
First citationUhl, W. & Layh, M. (1991). J. Organomet. Chem. 415, 181–190.  CSD CrossRef CAS Google Scholar
First citationUhl, W. & Roesky, H. W. (2002). Inorganic Chemistry Highlights, pp. 357–389.  Google Scholar
First citationUhl, W., Rösener, C., Stefaniak, C., Choy, T., Jasper-Peter, B., Kösters, J., Layh, M. & Hepp, A. (2012b). Z. Naturforsch. B: J. Chem. Sci. 67, 1081–1090.  Google Scholar
First citationUhl, W., Rösener, C., Layh, M. & Hepp, A. (2012a). Z. Anorg. Allg. Chem. 638, 1746–1754.  CSD CrossRef CAS Google Scholar
First citationWehmschulte, R. J., Grigsby, W. J., Schiemenz, B., Bartlett, R. A. & Power, P. P. (1996). Inorg. Chem. 35, 6694–6702.  CSD CrossRef PubMed CAS Google Scholar
First citationYanagisawa, T., Mizuhata, Y. & Tokitoh, N. (2018). Heteroat. Chem. 29, e21465.  CSD CrossRef 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