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

Journal logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206

Bond-length distributions for ions bonded to oxygen: metalloids and post-transition metals

CROSSMARK_Color_square_no_text.svg

aGeological Sciences, University of Manitoba, 125 Dysart Road, Winnipeg, Manitoba R3T 2N2, Canada
*Correspondence e-mail: umgagneo@myumanitoba.ca

Edited by R. Černý, University of Geneva, Switzerland (Received 14 September 2017; accepted 5 December 2017; online 12 January 2018)

Bond-length distributions have been examined for 33 configurations of the metalloid ions and 56 configurations of the post-transition metal ions bonded to oxygen, for 5279 coordination polyhedra and 21 761 bond distances for the metalloid ions, and 1821 coordination polyhedra and 10 723 bond distances for the post-transition metal ions. For the metalloid and post-transition elements with lone-pair electrons, the more common oxidation state between n versus n+2 is n for Sn, Te, Tl, Pb and Bi and n+2 for As and Sb. There is no correlation between bond-valence sum and coordination number for cations with stereoactive lone-pair electrons when including secondary bonds, and both intermediate states of lone-pair stereoactivity and inert lone pairs may occur for any coordination number > [4]. Variations in mean bond length are ∼0.06–0.09 Å for strongly bonded oxyanions of metalloid and post-transition metal ions, and ∼0.1–0.3 Å for ions showing lone-pair stereoactivity. Bond-length distortion is confirmed to be a leading cause of variation in mean bond lengths for ions with stereoactive lone-pair electrons. For strongly bonded cations (i.e. oxyanions), the causes of mean bond-length variation are unclear; the most plausible cause of mean bond-length variation for these ions is the effect of structure type, i.e. stress resulting from the inability of a structure to adopt its characteristic a priori bond lengths.

1. Introduction

This paper is the third in a series [Gagné & Hawthorne (2016a[Gagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602-625.], 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]); see also Gagné (2018[Gagné, O. C. (2018). Acta Cryst. B74, 49-62.]) in this issue] on the bond-length distributions of ions bonded to oxygen in crystals, and will focus on the metalloid and post-transition metal ions. For a detailed introduction and rationale for this work and a description of the data-collection and data-filtering methods, see Gagné & Hawthorne (2016a[Gagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602-625.]). In this series, we examine the distribution of bond lengths for 135 ions bonded to oxygen in 462 configurations using 180 331 bond lengths extracted from 9367 refined crystal structures; these data involve most ions of the periodic table and all coordination numbers in which they occur. Working with a large amount of data allows examination of subtle differences between the shapes of various distributions (e.g. bond-length distributions, mean bond-length distributions) which reflect differences in their structural and/or electronic behaviour. The factors that affect bond lengths are of general interest to all who work on crystal structures and their properties, and a comprehensive analysis of all the data should lead to increased understanding of those factors. Moreover, knowledge of possible variation in bond lengths is important in evaluating computational results on structural arrangements by setting expectations and limits as to what bond lengths may be observed between ion pairs, and are also useful in identifying unusual stereochemical features in new crystal structures.

Here, we report the data and bond-length distributions for nine metalloid ions and 11 post-transition metal ions bonded to O2−: we report 33 configurations of the metalloid ions as a function of coordination number when bonded to O2− (21 761 bond lengths and 5279 coordination polyhedra from 2575 crystal structure refinements), and 56 configurations for the post-transition metals (10 723 bond lengths and 1821 coordination polyhedra from 1143 crystal structure refinements). This article covers some strongly bonded oxyanions (e.g. BO3, SiO4) and ions with stereoactive lone-pair electrons (e.g. Sn2+, Tl+), and complements our discussion of these types of ions for non-metals bonded to O2− (Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]).

2. Lone-pair stereoactivity

Of the 135 ions for which we have gathered data in our bond-length dispersion analysis, we observe 14 cations with lone-pair electrons bonded to O2−, and 11 ions with stereoactive lone-pair electrons bonded to O2−. For the ions with stereoactive lone-pair electrons, seven ions are non-metals, three ions are metalloids and four ions are post-transition metals. For a thorough discussion of lone-pair stereoactivity and a general analysis for the 11 ions with stereoactive lone-pair electrons bonded to O2−, we refer the reader to the second paper of this series (Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]); here, we reiterate some important points, and give a more detailed discussion of lone-pair stereoactivity for the metalloids and post-transition metals later in text.

Lone-pair stereoactivity is associated with the ns2np0 electron configuration of p-block cations and the formation of highly anisotropic coordination polyhedra. Lone-pair stereoactive ions typically form short (strong) bonds in one hemisphere of their coordination shell, and long (weak) bonds in the other; these are commonly called `primary' and `secondary' bonds (Alcock, 1972[Alcock, N. W. (1972). Adv. Inorg. Radiat. Chem. 15, 1-57.]). Lone-pair stereoactivity has successfully been explained via atomic orbital arguments (see below), but we note that it has also been rationalized using strictly Lewis acid–base arguments by Brown & Faggiani (1980[Brown, I. D. & Faggiani, R. (1980). Acta Cryst. B36, 1802-1806.]) and Brown (1988[Brown, I. D. (1988). Acta Cryst. B44, 545-553.], 2011[Brown, I. D. (2011). J. Phys. Chem. A, 115, 12638-12645.]) with some success.

Orgel (1959[Orgel, L. E. (1959). J. Chem. Soc. pp. 3815-3819.]) first described the origins of the stereochemical behaviour for ns2np0 cations based on the mixing of the non-bonding s and p orbitals of these cations in non-cubic environments. Orgel argued that the sp-hybridized orbitals, where the stereoactive lone-pair electrons reside, can only form at non-centrosymmetric sites due to the parity constraint of these orbitals, and that this can only be achieved via large distortions of the coordination polyhedra. Bersuker (1984[Bersuker, I. B. (1984). The Jahn-Teller Effect and Vibronic Interactions in Modern Chemistry. New York: Plenum.]) explained the occurrence of lone-pair stereoactivity via an energetically favourable interaction between the highest occupied molecular orbital (HOMO) of the cation and the lowest unoccupied molecular orbital (LUMO) of the anion; this was supported by many investigations in the years following (e.g. Lefebvre et al., 1987[Lefebvre, I., Lannoo, M., Allan, G., Ibanez, A., Fourcade, J. & Jumas, J. C. (1987). Phys. Rev. Lett. 59, 2471-2474.], 1998[Lefebvre, I., Szymanski, M. A., Olivier-Fourcade, J. & Jumas, J. C. (1998). Phys. Rev. B, 58, 1896-1906.]; Watson & Parker, 1999[Watson, G. W. & Parker, S. C. (1999). J. Phys. Chem. B, 103, 1258-1262.]; Watson et al., 1999[Watson, G. W., Parker, S. C. & Kresse, G. (1999). Phys. Rev. B, 59, 8481-8486.]; Seshadri & Hill, 2001[Seshadri, R. & Hill, N. A. (2001). Chem. Mater. 13, 2892-2899.]; Waghmare et al., 2003[Waghmare, U. V., Spaldin, N. A., Kandpal, H. C. & Seshadri, R. (2003). Phys. Rev. B, 67, 125111.]; Stoltzfus et al., 2007[Stoltzfus, M. W., Woodward, P. M., Seshadri, R., Klepeis, J.-H. & Bursten, B. (2007). Inorg. Chem. 46, 3839-3850.]). From these findings, Walsh et al. (2011[Walsh, A., Payne, D. J., Egdell, R. G. & Watson, G. W. (2011). Chem. Soc. Rev. 40, 4455-4463.]) gave a revised model of lone-pair stereoactivity with explicit dependence on the anion, where strong interactions between the cation s and anion p orbitals result in a high-energy antibonding state, which, via distortion of the crystal structure, may interact with the empty cation p orbitals to form an electronic state where the lone pair resides.

We note that although the VSEPR model (Gillespie & Nyholm, 1957[Gillespie, R. J. & Nyholm, R. S. (1957). Quart. Rev. Chem. Soc. 11, 339-380.]; Gillespie, 1972[Gillespie, R. J. (1972). Molecular Geometry. London: Van Nostrand Reinhold.]) is commonly used to illustrate the bonding geometry of ions with stereoactive lone-pair electrons, it provides no driving mechanism for lone-pair stereoactivity/inactivity, and fails to explain the many cases for which lone-pair electrons are inactive, i.e. high-symmetry environments.

3. Coordination number

Whereas coordination number may be defined in simple terms, e.g. the number of counterions bonded to an ion, the decision to consider atom pairs as `bonded' is not obvious in many situations. This is particularly true for ions with stereoactive lone-pair electrons, as their coordination polyhedra are prone to large distortions, can form secondary bonds (up to ∼4 Å in length), and may be observed in a wide spectrum of `intermediate states' between stereoactivity and inactivity of the lone-pair electrons (Galy et al., 1975[Galy, J., Meunier, G., Andersson, S. & Åström, A. (1975). J. Solid State Chem. 13, 142-159.]).

Gagné & Hawthorne (2015[Gagné, O. C. & Hawthorne, F. C. (2015). Acta Cryst. B71, 562-578.], 2016a[Gagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602-625.]) provided arguments for including the longer interatomic distances of the first coordination shell for lone-pair stereoactive ions and for the larger alkali and alkaline earth metals as `bonded', by analyzing (1) trends in the bond-valence parameters of these ions, and (2) the gap between the first and second coordination shell. This analysis is summarized in the previous paper of this series (Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]). As we did for the non-metal ions with lone-pair stereoactive electrons, here we derive coordination polyhedra using the method described by Gagné & Hawthorne (2016a[Gagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602-625.]), which leads to the inclusion of all interatomic distances in the first coordination shell of the cations. This method leads to observed coordination numbers up to [12] for four lone-pair stereoactive cations, Tl+, Pb2+, Bi3+ and Te4+, and to coordination numbers up to [14] for Ba2+, [15] for K+, [18] for Rb+, and [20] for Cs+ (Gagné & Hawthorne, 2016a[Gagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602-625.]).

The inclusion of the `longer interatomic distances' follows the work of Alig & Trömel (1992[Alig, H. & Trömel, M. (1992). Z. Kristallogr. 201, 213-222.]) as well as that of Preiser et al. (1999[Preiser, C., Lösel, J., Brown, I. D., Kunz, M. & Skowron, A. (1999). Acta Cryst. B55, 698-711.]) who provided theoretical evidence that some of the longer cation–anion distances (up to 4 Å) may contribute to weak but significant chemical bonding via the calculation of electrostatic fluxes.

4. Sample size

Dealing with a very large amount of data has allowed us to critically evaluate the reproducibility of our results as a function of sampling. We described the effects of sample size (e.g. the presence of outliers, non-random sampling) in the first paper of this series (Gagné & Hawthorne, 2016a[Gagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602-625.]), as well as the effect of sample size on grand mean bond length (and its standard deviation), skewness, and kurtosis for [6]Na+ bonded to O2−. We reported the effect of sample size on these values for [4]S6+ and [6]I5+ bonded to O2− in the second paper of this series (Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]). Here, we do a similar analysis for [4]Si4+ and for [8]Bi3+. This analysis is done to sample bond strengths not covered by Gagné & Hawthorne (2016a[Gagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602-625.], 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]), as Gagné & Hawthorne (2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]) showed dependence of grand mean bond length, skewness and kurtosis values on bond strength and multi-modality of the bond-length distribution. Here we sample similar but weaker bonds for Si—O (mean bond valence 1 v.u.) compared to [4]S6+—O2− (mean bond valence 1.5 v.u.), and for [8]Bi3+—O2− (0.375 v.u.) compared with [6]I5+—O2− (0.83 v.u.) for lone-pair stereoactive cations. We report the sample sizes as a function of the number of coordination polyhedra.

Fig. 1[link] shows that for [4]Si4+, variation of less than ±0.005 Å in grand mean bond length is observed for sample sizes greater than 25 coordination polyhedra, while reliable values of skewness (±0.2) and kurtosis (±0.6) are obtained for sample sizes greater than 70 coordination polyhedra. For [8]Bi3+ (Fig. 2[link]), variability of less than ±0.005 Å is observed for 70 or more coordination polyhedra. However, it is possible that an appropriate sample size requires more than 70 coordination polyhedra but is limited here by the size of the parent distribution. Reliable values of skewness and kurtosis are obtained for sample sizes greater than seven coordination polyhedra.

[Figure 1]
Figure 1
The effect of sample size on (a) mean bond length, (b) standard deviation of the mean bond length, (c) skewness, and (d) kurtosis for [4]Si4+. The dashed line shows the value for the parent distribution.
[Figure 2]
Figure 2
The effect of sample size on (a) mean bond length, (b) standard deviation of the mean bond length, (c) skewness, and (d) kurtosis for [8]Bi3+. The dashed line shows the value for the parent distribution.

Thus [8]Bi3+ compares very well with [6]I5+ (∼40 coordination polyhedra for the same level of agreement for grand mean bond lengths, and only two coordination polyhedra for skewness and kurtosis; Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]) despite significantly weaker bond strengths, due to the overwhelming effect of lone-pair stereoactivity on the bond-length distributions of these ions. For [4]Si4+, more data is needed than for [4]S6+ (approximately five coordination polyhedra; Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]) for a reliable value of the grand mean bond length, probably due to the formation of relatively weaker bonds. However, significantly less data are needed for [4]Si4+ in comparison to [4]S6+ (∼300 coordination polyhedra; Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]) to obtain reliable values of skewness and kurtosis.

Mean bond-length distributions were analyzed in a similar way. Minimum sample sizes were determined for the skewness and kurtosis of these distributions with the same cut-offs as above, less than which these values have little significance and are not reported. For [4]Si4+, the threshold was observed at ∼400 coordination polyhedra (∼700 for [4]S6+; Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]) and ∼60 coordination polyhedra for [8]Bi3+, (∼50 for [6]I5+; Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]).

5. Results

5.1. Metalloids

For the metalloid ions bonded to O2−, the collection and filtering criteria described in Gagné & Hawthorne (2016a[Gagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602-625.]) resulted in a sample of 21 761 bonds and 5279 coordination polyhedra. Table 1[link] gives the mean bond length and standard deviation, the minimum and maximum bond length (and range), the skewness and kurtosis (where justified by sample size), and the number of bonds and coordination polyhedra for the 33 configurations for which the nine metalloid ions are observed in. All bond-length and bond-valence distributions (using the bond-valence parameters of Gagné & Hawthorne, 2015[Gagné, O. C. & Hawthorne, F. C. (2015). Acta Cryst. B71, 562-578.]) are shown in Figs. S1 and S2 (supporting information), respectively; bond-length distributions with adequate sample sizes (see above) are given in Fig. 3[link].

Table 1
Bond-length statistics for the metalloid ions bonded to O2−

Ion Coordination number Number of bonds Number of coordination polyhedra Mean bond length (Å) Standard deviation (Å) Range (Å) Maximum bond length (Å) Minimum bond length (Å) Skewness Kurtosis
B3+ 3 2556 852 1.372 0.021 0.166 1.464 1.298 0.2 1.0
4 2688 672 1.475 0.028 0.236 1.616 1.380 0.6 1.4
Si4+ 4 9128 2282 1.625 0.024 0.166 1.726 1.560 0.0 0.0
6 144 24 1.783 0.028 0.197 1.903 1.706
Ge4+ 4 1116 279 1.752 0.027 0.179 1.859 1.680 0.4 0.6
5 60 12 1.847 0.080 0.398 2.117 1.719
6 264 44 1.894 0.033 0.177 1.995 1.818
As3+ 3 27 9 1.776 0.045 0.174 1.845 1.671
4 16 4 2.026 0.406 1.177 2.896 1.719
5 40 8 2.247 0.553 1.338 3.092 1.754 0.5 −1.8
6 36 6 2.410 0.620 1.492 3.212 1.720 0.1 −2.0
8 8 1 2.480 0.507 1.201 3.005 1.804
As5+ 4 2032 508 1.687 0.027 0.196 1.806 1.610 0.7 1.5
6 78 13 1.830 0.028 0.121 1.888 1.767
Sb3+ 3 18 6 1.932 0.024 0.083 1.982 1.899
  4 60 15 2.092 0.178 0.698 2.596 1.898 1.3 1.0
  5 35 7 2.240 0.309 1.018 2.963 1.945 1.2 0.4
  6 72 12 2.443 0.485 1.474 3.391 1.917 0.4 −1.5
  7 49 7 2.486 0.451 1.308 3.266 1.958 0.3 −1.5
  8 32 4 2.584 0.477 1.386 3.362 1.976
  9 9 1 2.758 0.583 1.404 3.385 1.981
Sb5+ 6 948 158 1.977 0.034 0.208 2.102 1.894 0.7 1.0
Te4+ 3 12 4 1.843 0.015 0.043 1.862 1.819
  4 28 7 1.984 0.123 0.329 2.176 1.847 0.2 −1.9
  5 80 16 2.251 0.511 1.773 3.586 1.813 1.2 0.0
  6 360 60 2.386 0.552 1.786 3.595 1.809 0.5 −1.3
  7 245 35 2.460 0.538 1.747 3.556 1.809 0.2 −1.6
  8 512 64 2.594 0.608 1.805 3.615 1.810 0.1 −1.6
  9 126 14 2.677 0.611 1.830 3.674 1.844 −0.2 −1.5
  10 60 6 2.833 0.692 1.873 3.736 1.863 −0.3 −1.6
  11 22 2 2.812 0.668 1.699 3.581 1.882
  12 36 3 2.928 0.643 1.853 3.697 1.844
Te6+ 6 864 144 1.923 0.030 0.231 2.048 1.817 0.7 3.5
[Figure 3]
Figure 3
Bond-length distributions for selected configurations of the metalloid ions bonded to O2−: (a) [3]B3+, (b) [4]B3+, (c) [4]Si4+, (d) [6]Si4+, (e) [4]Ge4+, (f) [6]Ge4+, (g) [4]As5+, (h) [6]As5+, (i) [6]Sb3+, (j) [6]Sb5+, (k) [5]Te4+, (l) [6]Te4+, (m) [7]Te4+, (n) [8]Te4+, (o) [9]Te4+, (p) [6]Te6+.

An important issue in proposing bond-length ranges that ions may adopt is the reliability of the data at the limits of its distribution, i.e. the shortest and longest bonds of each ion configuration; below we examine some of the data at the lower and upper limits of these distributions. Special attention was paid to identifying short and long bond lengths that were the result of disorder, substitution of other ions, anomalous displacement parameters and uncorrected twinning effects.

5.1.1. B3+

B3+ occurs in three coordination numbers [2], [3] and [4], with a slight preference for [3] relative to [4]. [2]-coordination occurs in only four structures (Calvo & Faggiani, 1974[Calvo, C. & Faggiani, R. (1974). J. Chem. Soc. Chem. Commun. pp. 714-715.]; Calvo et al., 1975[Calvo, C., Faggiani, R. & Krishnamachari, N. (1975). Acta Cryst. B31, 188-192.]) where a BO2 group is aligned parallel to the c axis in synthetic apatite structures. The presence of B in these crystals was confirmed by chemical analysis and by 11B NMR. These structures did not quite pass our filters, but the occurrence of [2]B3+ was thought to be sufficiently significant that it should be noted.

[3]B3+—O2− distances are in the range 1.298–1.464 Å with a grand mean value of 1.372 Å; the latter is close to the value given for [3]B3+—O2− distances in minerals by Hawthorne et al. (1996[Hawthorne, F. C., Burns, P. C. & Grice, J. D. (1996). Rev. Mineral. Geochem. 33, 41-115.]): 1.370 Å. There is one very short [3]B3+—O2− distance of 1.298 Å in the structure of CsBO2 (Schläger & Hoppe, 1994[Schläger, M. & Hoppe, R. (1994). Z. Anorg. Allg. Chem. 620, 1867-1871.]). The constituent anion is coordinated by one [3]B3+ and five Cs+ anions between 3.029 and 3.251 Å with an incident bond-valence sum of 1.864 v.u.; this sum is low, although not unusually so, and would need an even shorter [3]B3+—O2− distance to increase the sum. Hence this value seems a reliable minimum distance at present. Er2Cl2(B2O5) (Nikelski & Schleid, 2003[Nikelski, T. & Schleid, T. (2003). Z. Anorg. Allg. Chem. 629, 2028-2033.]) has a [3]B3+—O2− distance of 1.453 Å to a O2− ion that bridges two (BO3) groups. The anion also bonds to Er3+ and the sum of the incident bond valences is 1.956 v.u. This is the longest reliable [3]B3+—O2− distance. The skewness of the distribution is very low, as expected for an ion with high bond valences and small coordination number.

[4]B3+—O2− distances are in the range 1.380–1.616 Å with a grand mean value of 1.475 Å; the value given for the grand mean [4]B3+—O2− distance in minerals by Hawthorne et al. (1996[Hawthorne, F. C., Burns, P. C. & Grice, J. D. (1996). Rev. Mineral. Geochem. 33, 41-115.]) is 1.476 Å. The structure of Gd2(B4O9) (Emme & Huppertz, 2003[Emme, H. & Huppertz, H. (2003). Chem. Eur. J. 9, 3623-3633.]) has edge-sharing (BO4) groups and both very short (1.380 Å) and very long (1.603 Å) [4]B3+—O2− distances. The sum of the incident bond valences at the central B3+ ion is 2.850 v.u and at the anions is 1.796 and 2.206 v.u. The structure is well refined and these distances seem reliable. Longer [4]B3+—O2− distances have been published: the structure of piergorite-(Ce) (Boiocchi et al., 2006[Boiocchi, M., Callegari, C. & Ottolini, L. (2006). Am. Mineral. 91, 1170-1177.]) lists a [4]B3+—O2− distance of 1.664 (6) Å. However, the 〈[3]B3+—O2−〉 distance is 1.525 Å, much too large for occupancy of the tetrahedrally coordinated site by B3+ alone. In accord with this, the Ueq value for the central B site is 2.4× the mean value of the other three B sites in the structure, indicating that there is substitution for B3+ by a heavier cation, almost certainly Si4+ that leads to the anomalously large 〈[3]B3+—O2−〉 distance. Thus the data for piergorite-(Ce) was not included in our analysis.

5.1.2. Si4+

Si4+ occurs in two coordination numbers: [4] and [6], with a very strong preference for [4] over [6]: 2282 versus 24 polyhedra, respectively (Table 1[link]). [6]-coordination is generally associated with high-pressure phases, although thaumasite, Ca3Si(OH)6(CO3)(SO4)(H2O)12 (Jacobsen et al., 2003[Jacobsen, S. D., Smyth, J. R. & Swope, R. J. (2003). Phys. Chem. Mineral. 30, 321-329.]) contains [6]-coordinate Si4+ and occurs as a low-temperature secondary alteration phase in mafic igneous and metamorphic rocks. [4]Si4+—O distances are in the range 1.560–1.726 Å with a grand mean value of 1.625 Å, close to the value of 1.624 Å given by Baur (1978[Baur, W. H. (1978). Acta Cryst. B34, 1751-1756.]). Si4+—O distances smaller than 1.56 Å are commonly recorded, but are associated with high variability in Ueq values, substitution of B3+ and P5+ for Si4+, and/or disorder of other cations in the structure. Our estimate of a reliable minimum Si4+—O2− distance is 1.560 Å. In the type-B (high-pressure) R2Si2O7 (R = Gd, Tb, Dy, Ho) structures, the longest Si4+—O2− distances to bridging anions are in the range 1.708–1.725 Å in well refined structures (Fleet & Liu, 2003[Fleet, M. E. & Liu, X. (2003). Z. Kristallogr. 218, 795-801.]). In these four structures, the bond-valence sums at the anions involved in the longest Si4+—O2− distances are 2.09, 2.12, 2.09 and 2.12 v.u. for R3+ = Gd, Tb, Dy, Ho, respectively. Thus the longest reliable Si4+—O2− distance is 1.726 Å. The bond-length distribution for Si4+O4 has unusually low values of skewness (0.0) and kurtosis (0.0).

[6]Si4+—O2− distances are in the range 1.706–1.903 Å with a grand mean value of 1.783 Å. There is one very short [6]Si4+—O2− distance of 1.706 Å in the structure of SiPO4(OH) (Stearns et al., 2005[Stearns, L. A., Groy, T. L. & Leinenweber, K. (2005). J. Solid State Chem. 178, 2594-2601.]). However, the sum of the bond valences at the constituent anion is 2.077 v.u., suggesting that this is a valid distance. Pacalo & Parise (1992[Pacalo, R. E. G. & Parise, J. B. (1992). Am. Mineral. 77, 681-684.]) report a [6]Si4+—O2− distance of 1.903 Å, significantly larger that the next-lowest values around 1.83 Å. There is no apparent flaw in the structure refinement, but the sum of the bond valences incident at the constituent anion is 1.859 v.u. The [6]Si4+—O2− distance required for exact adherence to the valence-sum rule is 1.804 Å, within the range of values observed in other structures. However, the small number of data leave the possible maximum length of the [6]Si4+—O2− bond an open question.

5.1.3. Ge4+

Ge4+ occurs in three coordination numbers: [4], [5] and [6], with a very strong preference for [4] and a slight preference for [6] over [5]. [4]Ge4+—O2− distances are in the range 1.680–1.859 Å with a grand mean value of 1.752 Å. The largest value of 1.859 Å occurs in the structure of Ca5Ge3O11 (Barbier & Levy, 1997[Barbier, J. & Levy, D. (1997). Z. Kristallogr. 212, 519-528.]). The structure is well refined and the O2− ion bridges two (GeO4) tetrahedra and bonds to two additional Ca2+ ions for an incident bond-valence sum of 2.176 v.u. The [4]Ge4+—O distance of 1.844 Å occurs in the structure of Fe2Ge2O8 (Kato et al., 1979[Kato, K., Takayama, E. & Kimizuka, N. (1979). Naturwissenschaften, 66, 616-617.]); the constituent anion bridges two (GeO4) tetrahedra and bonds to an additional Fe2+ ion for an incident bond-valence sum of 2.046 v.u. Thus the tail to higher values in Fig. 3[link](e) is a result of a small number of linked (GeO4) tetrahedra in structures where the bridging anion bonds to other cations. Similar to [4]Si4+, the distribution of [4]Ge4+—O2− distances shows low skewness (0.4) and kurtosis (0.6) (Table 1[link]). [5]Ge4+—O distances are in the range 1.719–2.117 Å with a grand mean value of 1.847 Å, although the number of data is small. [6]Ge4+—O2− distances are in the range 1.818–1.995 Å with a grand mean value of 1.894 Å. The distribution of [6]Ge4+—O2− distances shows a tail to longer values, and examination of these structures shows that these distances involve O2− ions that bridge (GeO6) octahedra and link to other divalent cations.

5.1.4. As3+

As3+ occurs in five coordination numbers from [3] to [8] with an average observed coordination number of [5] and a grand mean bond length of 2.107 Å for 28 polyhedra. As3+ is strongly lone-pair stereoactive and despite the paucity of data, all coordination numbers above [3] show bimodal distributions of bond lengths. There are always three short primary bonds for all coordination numbers in the range 1.671–1.891 Å with a mean value of 1.793 Å, to be compared with a grand mean value of 1.783 Å for minerals given by Majzlan et al. (2014[Majzlan, J., Drahota, P. & Filippi, M. (2014). Rev. Mineral. Geochem. 79, 17-184.]). There is a gap of >0.80 Å between the primary bonds and the shortest secondary bonds for all coordination numbers > [3].

5.1.5. As5+

As5+ occurs in two coordination numbers: [4] and [6] with [4] dominant over [6] (Table 1[link]). For [4]-coordination, the grand mean bond length is 1.687 Å, close to the value of 1.685 Å given by Majzlan et al. (2014[Majzlan, J., Drahota, P. & Filippi, M. (2014). Rev. Mineral. Geochem. 79, 17-184.]) for minerals; the individual bond-length range is 1.610–1.806 Å. The distribution shows a long tail to larger values (Fig. 3[link]g), but these data are from well refined structures and are reasonable from a crystal-chemical point of view. The structure of CaK2As2O7 (Faggiani & Calvo, 1976[Faggiani, R. & Calvo, C. (1976). Can. J. Chem. 54, 3319-3324.]) has a diarsenate group; the bridging O2− ion has [4]As5+—O2− distances of 1.799 and 1.786 Å with an additional Ca2+—O2− bond of 2.875 Å for an incident bond-valence sum of 1.976 v.u. Distances of 1.795 and 1.790 Å to two different O2− ions are listed in the structure of TlH2AsO4 (Narasaiah et al., 1987[Narasaiah, T. V., Choudhary, R. N. P., Nigam, G. D. & Mattern, G. (1987). Acta Cryst. C43, 388-390.]). The constituent anions also bond to Tl+ at 2.949 and 2.965 Å for bond-valence sums of 1.043 and 1.049 v.u., respectively. In accord with the composition of the crystal, the valence-sum rule indicates that these anions are OH groups and each receives a bond-valence contribution from the associated H+ ion, bringing the incident bond-valence close to 2 v.u. Thus the tail of long values for [4]As5+—O2− bonds (Fig. 3[link]g) is due to a small number of polymerized and acid (AsO4) groups.

For [6]-coordination, the grand mean bond length is 1.830 Å with an individual bond-length range of 1.767–1.888 Å; this is the smallest range for any [6]-coordinated metalloid cation, although this may be a result of the small amount of data available (Table 1[link]).

5.1.6. Sb3+

Sb3+ occurs in seven coordination numbers from [3] to [9] with an average observed coordination number of [6] and a grand mean bond length of 2.278 Å for 52 polyhedra. Sb3+ is strongly lone-pair stereoactive. For [3]Sb3+, there are no secondary bonds and the grand mean bond length is correspondingly short: 1.932 with a range of 1.899–1.982 Å. The grand mean bond lengths increase monotonically with increasing coordination number as the number of secondary bonds increases. The number of primary bonds varies from three (most common) to two examples of five in NaSb3O2(PO4)2 (Adair et al., 2000[Adair, B. A., Diaz de Delgado, G., Delgado, J. M. & Cheetham, A. K. (2000). J. Solid State Chem. 151, 21-24.]): 2.310, 1.982, 2.301, 2.121, 2.149 Å and 2.038, 2.301, 2.147, 2.113, 2.296 Å, and the division between primary and secondary bonds is less pronounced than in other lone-pair stereoactive ions.

5.1.7. Sb5+

Sb5+ occurs in coordination number [6] with a grand mean bond length of 1.978 Å and a range of 1.894–2.102 Å for 183 polyhedra. The lower limit of 1.894 Å is indicated by several well refined structures with minimum Sb5+—O2− distances in the range 1.89–1.90 Å. The longest reliable Sb5+—O2− distance is in Sb2O5 (Jansen, 1978[Jansen, M. (1978). Angew. Chem. 90, 141-141.]) where an O2− ion is bonded to three Sb5+ ions at distances of 2.043, 2.085 and 2.102 Å for an incident bond-valence sum of 2.041 v.u., thus giving a crystal-chemical justification for the long observed Sb5+—O2− distance.

5.1.8. Te4+

Te4+ occurs in ten coordination numbers from [3] to [12] with most data observed in coordination numbers [6] and [8]; the grand mean bond length is 2.469 Å for 211 polyhedra. Te4+ is strongly lone-pair stereoactive and most of the coordination numbers show a bimodal distribution of bond lengths [Figs. 3(k)–3(o), and Figs. S1(x)–1(ag)]. For [3]Te4+, there are no secondary bonds and the grand mean bond length is correspondingly short: 1.843 with a range of 1.819–1.862 Å. The grand mean bond lengths increase monotonically with increasing coordination number as the number of secondary bonds increases. The number of primary bonds varies from three (most common) to five in NiTe2O5 (Platte & Trömel, 1981[Platte, C. & Trömel, M. (1981). Acta Cryst. B37, 1276-1278.]): 1.886, 1.996 ×2, 2.247 ×2 Å, and Te3SeO8 (Pico et al., 1986[Pico, C., Castro, A., Veiga, M. L., Gutierrez Puebla, E., Monge, M. A. & Ruiz Valero, C. (1986). J. Solid State Chem. 63, 172-178.]): 1.886, 2.021, 2.032, 2.218 ×2 Å. As observed for Sb3+, this behaviour is somewhat different to that of Se4+ which shows only three primary bonds irrespective of its coordination number (Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]). The ranges of bond lengths found are broadly compatible with those of Christy et al. (2016[Christy, A. G., Mills, S. J. & Kampf, A. R. (2016). Mineral. Mag. 80, 415-545.]).

5.1.9. Te6+

Te6+ occurs only in coordination number [6] with a grand mean bond length of 1.923 Å and a range of 1.817–2.048 Å for 155 polyhedra, compatible with the results of Christy et al. (2016[Christy, A. G., Mills, S. J. & Kampf, A. R. (2016). Mineral. Mag. 80, 415-545.]). Much of the data is concentrated in the centre of the range and there are long tails to each side of the distribution. In other examples of such distributions, it has been our experience that much of the data in such long tails to the distribution are the result of extensive (atomic or stacking) disorder or unresolved twinning in the structure, inadequate absorption corrections for heavily absorbing structures. However, for Te6+ the situation is somewhat different. Such problem structures still occur, but other structures in the tails of the distribution look well refined and the resulting stereochemistry appears reasonable, at least from a bond-valence perspective. The structure of Na2Te2O7 (Meier & Schleid, 2006[Meier, S. F. & Schleid, T. (2006). Z. Anorg. Allg. Chem. 632, 2151-2151.]) has a short [6]Te6+—O2− distance of 1.817 Å and is bonded to three Na+ ions for an incident bond-valence sum of 1.874 v.u.

5.2. Post-transition metals

For the post-transition metals ions bonded to O2−, the collection and filtering criteria described in Gagné & Hawthorne (2016a[Gagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602-625.]) resulted in a sample size of 10 723 bonds and 1821 coordination polyhedra. Table 2[link] gives the bond-length statistics for the 56 configurations for which the 11 post-transition metal ions are observed in. All bond-length and bond-valence distributions are shown in Figs. S3 and S4, respectively; bond-length distributions with adequate sample sizes are shown in Fig. 4[link].

Table 2
Bond-length statistics for the post-transition metal ions bonded to O2−

Ion Coordination number Number of bonds Number of coordination polyhedra Mean bond length (Å) Standard deviation (Å) Range (Å) Maximum bond length (Å) Minimum bond length (Å) Skewness Kurtosis
Al3+ 4 1224 306 1.746 0.022 0.142 1.827 1.685 0.2 0.3
5 155 31 1.842 0.047 0.190 1.938 1.748
6 2718 453 1.903 0.040 0.262 2.054 1.792 0.6 0.7
Ga3+ 4 532 133 1.842 0.027 0.174 1.948 1.774 0.5 0.4
5 90 18 1.910 0.076 0.483 2.254 1.771
6 396 66 1.978 0.040 0.237 2.130 1.893 0.8 1.7
In3+ 6 660 110 2.142 0.042 0.301 2.324 2.023 0.9 1.8
7 28 4 2.218 0.071 0.365 2.450 2.085
8 16 2 2.275 0.058 0.155 2.351 2.196
Sn2+ 3 63 21 2.094 0.031 0.158 2.162 2.004
4 24 6 2.281 0.225 0.818 2.881 2.063 1.3 1.1
5 30 6 2.469 0.402 1.166 3.228 2.062 0.6 −1.5
6 54 9 2.508 0.310 1.183 3.251 2.068 0.6 −0.7
7 14 2 2.690 0.488 1.345 3.407 2.062
8 40 5 2.706 0.423 1.212 3.327 2.115 0.1 −1.7
9 9 1 2.828 0.490 1.155 3.312 2.157
Sn4+ 4 16 4 1.956 0.009 0.035 1.970 1.935
6 192 32 2.054 0.024 0.134 2.130 1.996
7 7 1 2.115 0.027 0.070 2.142 2.072
Tl+ 3 9 3 2.517 0.051 0.157 2.569 2.412
  4 12 3 2.726 0.231 0.867 3.245 2.378
  5 15 3 2.840 0.296 0.864 3.356 2.492
  6 48 8 2.887 0.194 0.816 3.271 2.455 0.0 −0.6
  7 56 8 2.976 0.226 0.790 3.366 2.576 −0.1 −1.4
  8 96 12 2.988 0.192 0.804 3.363 2.559 0.2 −0.7
  9 99 11 2.991 0.192 0.846 3.404 2.558 0.2 −0.8
  10 110 11 3.102 0.220 0.979 3.626 2.647 0.4 −0.6
  11 55 5 3.134 0.258 0.998 3.674 2.676 0.4 −0.6
  12 108 9 3.195 0.285 1.290 4.012 2.722 0.7 −0.3
Tl3+ 6 36 6 2.228 0.076 0.407 2.481 2.074
7 7 1 2.336 0.157 0.504 2.624 2.120
8 16 2 2.378 0.138 0.618 2.804 2.186
Pb2+ 3 15 5 2.210 0.047 0.142 2.291 2.149 0.7 −1.1
  4 68 17 2.357 0.108 0.464 2.638 2.174 0.5 0.1
  5 130 26 2.482 0.212 0.843 2.990 2.147 0.4 −1.0
  6 192 32 2.581 0.237 0.984 3.153 2.169 0.2 −0.7
  7 322 46 2.637 0.205 1.089 3.214 2.125 0.1 −0.2
  8 592 74 2.697 0.210 1.048 3.280 2.232 0.4 −0.3
  9 360 40 2.750 0.217 0.950 3.231 2.281 0.2 −0.7
  10 140 14 2.789 0.212 0.970 3.261 2.291 0.1 −0.7
  11 44 4 2.838 0.205 0.767 3.168 2.401
  12 204 17 2.827 0.113 0.722 3.200 2.478 0.4 1.9
Pb4+ 4 8 2 2.056 0.022 0.076 2.074 1.998
5 5 1 2.147 0.100 0.276 2.324 2.048
6 54 9 2.169 0.038 0.168 2.237 2.069
Bi3+ 3 27 9 2.069 0.037 0.130 2.151 2.021
4 84 21 2.211 0.163 0.805 2.790 1.985 0.9 1.1
5 90 18 2.317 0.195 0.732 2.785 2.053 0.5 −0.7
6 222 37 2.398 0.229 0.904 2.981 2.077 0.6 −0.7
7 259 37 2.497 0.303 1.097 3.154 2.057 0.4 −1.2
8 664 83 2.522 0.252 1.069 3.165 2.096 0.5 −0.5
9 162 18 2.604 0.302 1.081 3.158 2.077 0.1 −1.2
10 70 7 2.670 0.313 1.093 3.219 2.126 0.0 −1.0
12 12 1 2.671 0.264 0.897 3.183 2.286
Bi5+ 4 4 1 1.979 0.015 0.038 2.002 1.964
6 60 10 2.110 0.033 0.165 2.174 2.009
[Figure 4]
Figure 4
Bond-length distributions for selected configurations of the post-transition metal ions bonded to O2−: (a) [4]Al3+, (b) [5]Al3+, (c) [6]Al3+, (d) [4]Ga3+, (e) [5]Ga3+, (f) [6]Ga3+, (g) [6]In3+, (h) [3]Sn2+, (i) [6]Sn2+, (j) [6]Sn4+, (k) [8]Tl+, (l) [9]Tl+, (m) [10]Tl+, (n) [12]Tl+, (o) [4]Pb2+, (p) [5]Pb2+, (q) [6]Pb2+, (r) [7]Pb2+, (s) [8]Pb2+, (t) [9]Pb2+, (u) [10]Pb2+, (v) [12]Pb2+, (w) [4]Bi3+, (x) [5]Bi3+, (y) [6]Bi3+, (z) [7]Bi3+, (aa) [8]Bi3+, (ab) [9]Bi3+, (ac) [6]Bi5+.
5.2.1. Al3+

Al3+ has three coordination numbers: [4], [5] and [6]; [6] is dominant (n = 453) and then [4] (n = 306) with [5] being less common (n = 31). [4]Al3+ has a grand mean bond length of 1.746 Å and a range of 1.685–1.833 Å. Distances below 1.68 Å occur in several structures but they all have potential for cation disorder and we consider them unreliable. [5]Al3+ has a grand mean bond length of 1.842 Å and a range of 1.748–1.938 Å. [6]Al3+ has grand mean bond length of 1.903 Å and a range of 1.792–2.054 Å. Values longer than 2.054 Å are given in the literature but these are associated with replacement of Al3+ by other ions. The longest confirmed bond distance of 2.054 Å occurs in the structure of BeAl6O10 (Alimpiev et al., 2002[Alimpiev, A. I., Merkulov, A. A., Solntsev, V. P., Tsvetkov, E. G., Matrosov, V. N. & Pestryakov, E. V. (2002). J. Cryst. Growth, 237, 884-889.]) in which the constituent O2− ion is coordinated by four Al3+ ions at distances of 2.054, 1.998, 1.880 and 1.861 Å for an incident bond-valence sum of 1.99 v.u.

5.2.2. Ga3+

Ga3+ has three coordination numbers: [4], [5] and [6]; [4] is dominant (n = 133) over [6] (n = 66) and [5] (n = 18). [4]Ga3+ has a grand mean bond length of 1.842 Å and a range of 1.774–1.948 Å. The minimum reliable distance occurs in BaGa2O4 (Kahlenberg et al., 2000[Kahlenberg, V., Fischer, R. X. & Parise, J. B. (2000). J. Solid State Chem. 154, 612-618.]) in which there is one O2− ion bridging two GaO4 tetrahedra and not bonded to Ba2+. The [4]Ga3+—O2− distances are 1.774 and 1.801 Å; the sum of the incident bond valences is low at 1.724 v.u. but this may be due to structural strain as BaGa2O4 was synthesized at high temperature (1350oC) and a [2]-coordinated bridging anion has little possibility of relaxation with decreasing temperature except shortening of its [4]Ga3+—O2− bonds. The longest [4]Ga3+—O2− bonds are in the range 1.91–1.94 Å and occur in the structures of Sr4(Ga2O7) (Kahlenberg et al., 2005[Kahlenberg, V., Lazic, B. & Krivovichev, S. V. (2005). J. Solid State Chem. 178, 1429-1439.]) and Ba4(Ga2O7) (Kahlenberg, 2001[Kahlenberg, V. (2001). Z. Anorg. Allg. Chem. 627, 2386-2390.]). [5]Ga has grand mean bond length of 1.910 Å and a range of 1.771–2.254 Å. The value of 2.254 Å is a very prominent outlier in the distribution of bond lengths [Fig. S3(e)]. It occurs in the structure of NaGa2(OH)(PO4)2 (Guesdon et al., 2003[Guesdon, A., Monnin, Y. & Raveau, B. (2003). J. Solid State Chem. 172, 237-242.]); the constituent O2− ion bonds to three Ga3+ ions and a H+ ion, and the incident bond-valence omitting the H+ ion is 1.06 v.u., suggesting that this long distance is valid. [6]Ga3+ has grand mean bond length of 1.978 Å and a range of 1.893–2.130 Å. The distribution shows a long tail to longer values.

5.2.3. In3+

In3+ has three coordination numbers: [6], [7] and [8]. Coordination number [6] is most frequently observed and has a grand mean bond length of 2.142 Å and a range of 2.023–2.324 Å. Many of the shortest and longest reliable In3+—O2− distances occur in the well ordered structure of In4(P2O7)3 (Thauern & Glaum, 2003[Thauern, H. & Glaum, R. (2003). Z. Anorg. Allg. Chem. 629, 479-486.]). Using the coordination numbers of the cations, there are 4 × 6 + 6 × 4 = 48 bonds in the structure, and hence the mean coordination number of the anions is 48/21 = [2.29]; 15 anions have a coordination number of [2] and six anions have a coordination number of [3]. The [2]-coordinated anions bond to In3+ and P5+ for a Pauling bond-strength sum of 1.75 v.u. and hence the bonds to these cations must be shorter than usual. Thus the In3+—O2− distances are very short, with three bonds in the range 2.023–2.036 Å, and the corresponding anion bond-valence sums are in the range 2.051–2.064 v.u. The [3]-coordinated anions bond to In3+ ×2 and P5+ for a Pauling bond-strength sum of 2.25 v.u. and hence the bonds to these cations must be longer than usual. As the structure contains P2O7 groups, the terminal P5+—O2− bonds cannot lengthen significantly and thus the reduction in incident bond-valence must be accommodated by elongation of the In3+—O2− bonds. Accordingly, there are six In3+—O2− bonds in the range 2.24–2.30 Å, accounting for many of the long In3+—O2− distances in Fig. 4[link](g), and the incident bond-valence sums are in the range 2.05–2.09 v.u. The longest reliable [6]In3+—O2− distance (2.324 Å) occurs in the structure of CuInW2O8 (Müller-Buschbaum & Szillat, 1994[Müller-Buschbaum, H. & Szillat, H. (1994). Z. Anorg. Allg. Chem. 620, 642-646.]). The constituent anion also bonds to Cu2+ and W5+/6+ with an incident bond valence of 1.974 v.u.

5.2.4. Sn2+

Sn2+ occurs in seven coordination numbers from [3] to [9] with a grand mean bond length of 2.336 Å for 50 polyhedra. Sn2+ is strongly lone-pair stereoactive. For [3]Sn2+, there are no secondary bonds and the grand mean bond length is 2.094 Å with a range of 2.004–2.162 Å. For [4]Sn2+, the fourth bond distance varies from 2.470 to 2.881 Å. In most cases, all four distances should be regarded as primary bonds as they lie to one side (i.e. in one hemisphere of space to one side) of the cation. Thus in Sn2(S2O4)2 (Magnusson & Johansson, 1982[Magnusson, A. & Johansson, L. G. (1982). Acta Chem. Scand. A36, 429-433.]), the four distances 2.236, 2.242, 2.264 and 2.324 Å lie to one side of the Sn2+ ion, whereas in Sn3O(OH)PO4 (Jordan et al., 1980[Jordan, T. H., Dickens, B., Schroeder, L. W. & Brown, W. E. (1980). Inorg. Chem. 19, 2551-2556.]), one Sn2+ has four primary bonds at 2.065, 2.167, 2.281 and 2.470 Å whereas another Sn2+ has three primary bonds at 2.111, 2.138 and 2.167 Å and one secondary bond at 2.674 Å. Thus both arrangements, three primary bonds and four primary bonds, occur for [4]Sn2+. The grand mean bond length increases monotonically with increasing coordination number as the number of secondary bonds increases (Table 2[link]).

5.2.5. Sn4+

Sn4+ occurs in three coordination numbers, [4], [6] and [7]. Coordination number [6] is observed most frequently, and has a grand mean bond length of 2.054 Å with a range of 1.996–2.130 Å. The range of distances for [4]Sn4+ is very small, 1.935–1.970 Å, with a mean value of 1.956 Å; this may be the result of the small number of data and the restricted range of compositions (alkali metal stannates).

5.2.6. Tl+

Tl+ occurs in ten coordination numbers from [3] to [12], all with a small number of data. [3]Tl+ is strongly lone-pair stereoactive with three short bonds to one side of the Tl+ ion, as in Tl6Si2O7 (Piffard et al., 1975[Piffard, Y., Marchand, R. & Tournoux, M. (1975). Rev. Chim. Minér. 12, 210-217.]). [4]Tl+ also occurs in the structure of Tl6Si2O7 and is also lone-pair stereoactive with all four anions occurring on one side of the [4]Tl+ ion at distances of 2.378, 2.721, 2.812 and 3.014 Å. There is a gradual increase in mean Tl+—O2− distances with increasing coordination number. The difference between the primary and secondary bond lengths is not as great for Tl+ as for other lone-pair stereoactive ions, a result of the lower formal charge of Tl+ relative to other ions such as As3+ or Te4+.

5.2.7. Tl3+

Tl3+ occurs in three coordination numbers: [6], [7] and [8], and there is very little data (Table 2[link]). The grand mean bond lengths increase from 2.228 to 2.336 to 2.378 Å with increasing coordination number.

5.2.8. Pb2+

Pb2+ occurs in ten coordination numbers from [3] to [12] with a preference for [8]. The grand mean bond length is 2.680 Å for 275 polyhedra. For [3]Pb2+, there are no secondary bonds and the grand mean bond length is 2.210 Å with a range of 2.149–2.291 Å and all bonds lying to one side of the Pb2+ ion. For [4]Pb2+, the fourth (long) distance varies from 2.367 to 2.638 Å, and in most cases, all four O2− anions lie to one side of the cation. With increasing coordination number, there is no obvious development of bimodal distributions of bond lengths [Figs. 4(o)–4(v)] with the possible exception of [5] (Fig. 4[link]p). For [5]-coordination, all bonds can still be to one side of the cation, as in PbAl2O4 (Ploetz & Müller-Buschbaum, 1982[Ploetz, K. B. & Müller-Buschbaum, H. (1982). Z. Anorg. Allg. Chem. 488, 38-44.]). With increasing coordination number, this asymmetric distribution of coordinating anions can be lost, as in Pb(WO4) (Richter et al., 1976[Richter, P. W., Kruger, G. J. & Pistorius, C. W. F. T. (1976). Acta Cryst. B32, 928-929.]) in which the eight bonds seem distributed reasonably randomly around the central cation.

5.2.9. Pb4+

Pb4+ occurs in coordinations [4], [5] and [6] with a preference for [6]. The grand mean bond lengths increase with increasing coordination number, but the paucity of data (Table 2[link]) prevents any general conclusions.

5.2.10. Bi3+

Bi3+ occurs in nine coordination numbers from [3] to [12] with a marked preference for coordination [8] (Table 2[link]). The grand mean bond length is 2.481 Å for 231 polyhedra. Bi3+ is strongly lone-pair stereoactive, but there is little sign of bimodal distributions of bond lengths [Figs. 4(w)–4(ab)], except perhaps for a coordination of [7] (Fig. 4[link]z). For [3]Bi3+, there are no secondary bonds and the grand mean bond length is 2.069 Å with a range of 2.002–2.151 Å. For [4]Bi3+, the fourth bond distance varies from 2.291–2.790 Å but is always a primary bond in that it lies within the hemisphere containing the primary (short) bonds. For [5]Bi3+, the fifth bond distance varies from 2.336 to 2.785 Å but again is always a primary bond. For [6]Bi3+, the fifth bond distance varies from 2.409 to 2.981 Å. The bonds do not now occupy a single hemisphere, but one or two just project into the second hemisphere; this is the case both for Bi(PO3)3 (Palkina & Jost, 1975[Palkina, K. & Jost, K.-H. (1975). Acta Cryst. B31, 2281-2285.]) with a fifth bond of 2.435 Å and uranosphaerite, Bi(UO2)O2OH (Hughes et al., 2003[Hughes, K. A., Burns, P. C. & Kolitsch, U. (2003). Can. Mineral. 41, 677-685.]) with a fifth bond of 2.981 Å. At higher coordination numbers, the bonds are distributed more uniformly around the central cation, but the shortest three bonds still tend to be concentrated to one side of the central cation.

5.2.11. Bi5+

Bi5+ occurs in coordinations [4] and [6] with a preference for [6]. The distribution of distances for [6]-coordination shows a strong negative skewness, but this is probably the result of insufficient data (ten coordination polyhedra), as most other ions with strong negative skewness are characterized by very few data. The grand mean bond length for [6]Bi5+ is 2.110 Å with a range of 2.009–2.174 Å.

6. Discussion

6.1. Lone-pair stereoactivity for metalloids and post-transition metals

In our bond-length dispersion analysis, three metalloid and four post-transition metal cations bonded to O2− display lone-pair stereoactivity. These ions also occur in an n+2 oxidation state, i.e. with no lone-pair electrons. For the metalloids, the number of coordination polyhedra for the n and n+2 oxidation states are 28 versus 526 for As3+ and As5+, 54 versus 183 for Sb3+ and Sb5+, and 212 versus 155 for Te4+ and Te6+. For the post-transition metals, these numbers are 50 versus 38 for Sn2+ and Sn4+, 74 versus nine for Tl+ and Tl3+, 276 versus 12 for Pb2+ and Pb4+, and 231 versus 11 for Bi3+ and Bi5+. Therefore, two of the seven ions are observed more often in their n+2 oxidation state: the group 15 ions of periods 4 and 5, As3+ and Sb3+. Although this is also the case for P (the group 15 cation of period 3), this trend does not extend to Bi in period 6. In contrast, Gagné & Hawthorne (2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]) showed that the period 3 non-metal ions more frequently occur in their highest oxidation state (without lone pair) when bonded to O2− (P5+, S6+ and Cl7+), and in their lowest oxidation state (with lone pair) for the period 4 and 5 non-metals bonded to O2− (Se4+, Br5+, I5+).

When they are bonded to O2−, the metalloid and post-transition metal ions with stereoactive lone-pair electrons show no trend for the bond-length range and the skewness and kurtosis of the bond-length distribution; this is probably due to small sample size. In Fig. 5[link](a), we give mean bond length as a function of coordination number for the seven ions for sample sizes greater than five coordination polyhedra. Individual data points are clearly prone to error due to small sample size, but we nonetheless observe a somewhat regular increase with coordination number for these ions. The mean bond length may appear to increase in a logarithmic way for certain ions, but this is again probably due to small sample size for higher coordination numbers. For the alkali and alkaline earth metals bonded to O2−, Gagné & Hawthorne (2016a[Gagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602-625.]) reported a linear increase for larger sample sizes. In Fig. 5[link](b), we give deviations from the bond-valence sum for the cations as a function of coordination number for the same ions. Contrary to what was observed for alkali and alkaline earth metal ions bonded to O2− (Gagné & Hawthorne, 2016a[Gagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602-625.]), here we see no correlation between bond-valence sum and coordination number. The bond-valence parameters used are those of Gagné & Hawthorne (2015[Gagné, O. C. & Hawthorne, F. C. (2015). Acta Cryst. B71, 562-578.]), which were derived with a coordination-based optimization factor to minimize deviations as a function of coordination number; these deviations are otherwise large for other published sets of bond-valence parameters.

[Figure 5]
Figure 5
Values of (a) mean bond length (Å) and (b) mean bond-valence sum using the parameters of Gagné & Hawthorne (2015[Gagné, O. C. & Hawthorne, F. C. (2015). Acta Cryst. B71, 562-578.]) for the different coordination numbers of the metalloid and post-transition metal ions.
6.1.1. When do we observe lone-pair stereoactivity?

In a general examination of lone-pair stereoactivity for 14 non-metal, metalloid and post-transition metal cations with lone-pair electrons bonded to O2−, Gagné & Hawthorne (2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]) confirmed the observation of Galy et al. (1975[Galy, J., Meunier, G., Andersson, S. & Åström, A. (1975). J. Solid State Chem. 13, 142-159.]) that in the majority of cases, the lone-pair of cations is observed in an `intermediate state' between stereoactivity and inertness. They also showed that interatomic distances may be included as secondary bonds in 1126 of 1321 coordination polyhedra surveyed (∼85%). Where the lone pair is `fully stereoactive', the next-nearest anions are usually observed at 2–3× the distance of the mean bond length for the short bonds, too far and weak to be considered secondary bonds. Gagné & Hawthorne (2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]) also showed that lone-pair stereoactivity (as measured by bond-length distortion) correlates very poorly to coordination number for Se4+ and Pb2+ (R2 = 0.19 and 0.08, respectively), concluding that both intermediate and inert lone-pair electrons may occur for coordination numbers > [4]. In Fig. 6[link], we give a similar plot for (a) Bi3+ (p-value = 0.030, R2 = 0.02) and (b) Te4+ (p-value = 0.139, R2 = 0.01), confirming that there is no relation between lone-pair stereoactivity and coordination number for coordination numbers > [4].

[Figure 6]
Figure 6
Bond-length distortion as a function of coordination number for (a) Bi3+ and (b) Te4+. The p-values are (a) 0.030 and (b) 0.139.

This result follows the current model for lone-pair stereoactivity, which does not concern itself with coordination number. In this model (Walsh et al., 2011[Walsh, A., Payne, D. J., Egdell, R. G. & Watson, G. W. (2011). Chem. Soc. Rev. 40, 4455-4463.]), stereoactivity of the lone-pair electrons results from strong interactions between the cation s and anion p orbitals that result in a high-energy antibonding state. This antibonding state may then interact with the empty p orbitals of the cation via distortion of the crystal structure to form an electronic state where the lone pair resides; what coordination number will result from this is irrelevant to this phenomenon, and depends on the rest of the structure. Furthermore, whether or not distortion will result in a net stabilization of the occupied electronic states depends on the relative energy of the cation s and p and anion p orbitals, the prediction of which requires orbital energy calculations on a case-by-case basis.

Alternatively, Brown & Faggiani (1980[Brown, I. D. & Faggiani, R. (1980). Acta Cryst. B36, 1802-1806.]) showed that simple Lewis acid–base arguments may be used to predict lone-pair stereoactivity. They gave a loose inverse relation between the coordination number of Tl+ and the base strength of the anion, proposing that lone-pair electrons are always stereoactive where the counterion is a strong base with Lewis basicity > 0.22 v.u. However, a mixture of lone-pair stereoactivity and inactivity is observed below that threshold. Brown (1988[Brown, I. D. (1988). Acta Cryst. B44, 545-553.]) correlated a vector-based measure of bond-length distortion to the Lewis base strength of the anion for Tl+ structures, and updated the threshold to 0.27 v.u. This threshold is set to include as many structures with lone-pair stereoactive cations without including structures where the lone pair is inert. Structures observed above that threshold typically do not form secondary bonds and have coordination numbers [3] and [4]. Although the model may not be used to predict lone-pair stereoactivity below the set threshold (most cases), it may be used to predict lone-pair stereoactivity above it, i.e. for structure with strong anion complexes.

Thus the Lewis acid–base argument is easy to apply, but is not always useful. Although the procedure is more involved, the occurrence of lone-pair stereoactivity is more confidently predicted via orbital energy calculations.

6.2. Mean bond-length distributions

The mean bond-length distributions for the metalloid and post-transition metal ions bonded to O2− are given in Figs. S5 and S6, respectively. Those with adequate sample sizes (see sample size study above) are given in Figs. 7[link] and 8[link], and Tables 3[link] and 4[link] give the grand mean bond length (and standard deviation), the minimum and maximum mean bond length (and range), the skewness and kurtosis of these distributions (where justified by sample size) and the number of coordination polyhedra for all configurations observed.

Table 3
Mean bond-length statistics for the metalloid ions bonded to O2−

Ion Coordination number Number of coordination polyhedra Grand mean bond length (Å) Standard deviation (Å) Mean bond-length range (Å) Maximum mean bond length (Å) Minimum mean bond length (Å) Skewness Kurtosis
B3+ 3 852 1.372 0.008 0.062 1.401 1.339 0.2 0.5
4 672 1.475 0.008 0.073 1.517 1.444 0.9 3.9
Si4+ 4 2282 1.625 0.011 0.068 1.658 1.590 −0.1 0.0
6 24 1.783 0.017 0.064 1.819 1.755
Ge4+ 4 279 1.752 0.013 0.079 1.791 1.712 0.3 0.4
5 12 1.847 0.013 0.040 1.868 1.828
6 44 1.894 0.018 0.074 1.931 1.857
As3+ 3 9 1.776 0.011 0.036 1.794 1.758
4 4 2.026 0.036 0.081 2.077 1.996
5 8 2.247 0.025 0.082 2.295 2.213
6 6 2.410 0.060 0.163 2.495 2.332
8 1 2.480 0.000 2.480 2.480
As5+ 4 508 1.687 0.008 0.066 1.721 1.655 0.2 2.4
6 13 1.830 0.008 0.031 1.849 1.819
Sb3+ 3 6 1.932 0.022 0.060 1.974 1.914
4 15 2.092 0.025 0.064 2.125 2.061
5 7 2.240 0.063 0.182 2.361 2.179
6 12 2.443 0.083 0.274 2.623 2.349
7 7 2.486 0.023 0.072 2.517 2.445
8 4 2.584 0.032 0.074 2.626 2.552
9 1 2.758 0.000 2.758 2.758
Sb5+ 6 158 1.977 0.016 0.107 2.042 1.935
Te4+ 3 4 1.843 0.017 0.039 1.858 1.819
  4 7 1.984 0.011 0.042 2.006 1.964
  5 16 2.251 0.085 0.271 2.383 2.112
  6 60 2.386 0.096 0.573 2.660 2.087
  7 35 2.460 0.077 0.380 2.715 2.336
  8 64 2.594 0.082 0.323 2.776 2.453
  9 14 2.677 0.073 0.250 2.778 2.528
  10 6 2.833 0.085 0.203 2.921 2.718
  11 2 2.812 0.032 0.045 2.835 2.790
  12 3 2.928 0.102 0.193 3.044 2.851
Te6+ 6 144 1.923 0.013 0.081 1.971 1.890

Table 4
Mean bond-length statistics for the post-transition metal ions bonded to O2−

Ion Coordination number Number of coordination polyhedra Grand mean bond length (Å) Standard deviation (Å) Mean bond-length range (Å) Maximum mean bond length (Å) Minimum mean bond length (Å) Skewness Kurtosis
Al3+ 4 306 1.746 0.013 0.092 1.806 1.714 0.9 1.7
5 31 1.842 0.012 0.041 1.861 1.820
6 453 1.903 0.014 0.078 1.944 1.866 0.2 −0.1
Ga3+ 4 133 1.842 0.015 0.070 1.874 1.804
5 18 1.910 0.014 0.048 1.941 1.893
6 66 1.978 0.017 0.077 2.012 1.935
In3+ 6 110 2.142 0.018 0.101 2.198 2.097
7 4 2.218 0.014 0.030 2.238 2.208
8 2 2.275 0.002 0.003 2.276 2.274
Sn2+ 3 21 2.094 0.017 0.062 2.121 2.059
4 6 2.281 0.024 0.069 2.315 2.246
5 6 2.451 0.049 0.125 2.492 2.367
6 9 2.508 0.064 0.174 2.577 2.403
7 2 2.690 0.015 0.021 2.701 2.680
8 5 2.706 0.013 0.030 2.722 2.692
9 1 2.828 0.000 2.828 2.828
4 4 1.956 0.002 0.004 1.958 1.954
Sn4+ 6 32 2.054 0.010 0.044 2.077 2.033
7 1 2.115 0.000 2.115 2.115
Tl+ 3 3 2.517 0.035 0.070 2.551 2.481
  4 3 2.726 0.049 0.097 2.773 2.676
  5 3 2.840 0.055 0.108 2.889 2.781
  6 8 2.887 0.032 0.089 2.922 2.833
  7 8 2.976 0.030 0.094 3.012 2.918
  8 12 2.976 0.030 0.094 3.012 2.918
  9 11 2.991 0.030 0.095 3.049 2.954
  10 11 3.102 0.059 0.166 3.203 3.037
  11 5 3.134 0.043 0.113 3.176 3.062
  12 9 3.195 0.073 0.210 3.304 3.094
Tl3+ 6 6 2.228 0.013 0.036 2.242 2.206
7 1 2.336 0.000 2.336 2.336
8 2 2.378 0.047 0.067 2.412 2.345
Pb2+ 3 5 2.210 0.038 0.083 2.250 2.167
  4 17 2.357 0.030 0.119 2.413 2.294
  5 26 2.482 0.033 0.104 2.542 2.438
  6 32 2.581 0.037 0.195 2.665 2.470
  7 46 2.637 0.033 0.149 2.715 2.566
  8 74 2.697 0.036 0.170 2.778 2.608 0.2 0.5
  9 40 2.750 0.038 0.144 2.827 2.683
  10 14 2.789 0.028 0.100 2.840 2.740
  11 4 2.821 0.046 0.111 2.869 2.758
  12 17 2.827 0.030 0.094 2.867 2.773
Pb4+ 4 2 2.056 0.007 0.011 2.061 2.051
5 1 2.147 0.000 2.147 2.147
6 9 2.169 0.014 0.044 2.193 2.149
Bi3+ 3 9 2.069 0.021 0.060 2.096 2.036
4 21 2.211 0.024 0.095 2.268 2.173
5 18 2.317 0.033 0.098 2.369 2.271
6 37 2.398 0.031 0.149 2.485 2.336
7 37 2.497 0.043 0.193 2.580 2.387
8 83 2.522 0.052 0.230 2.639 2.409 0.0 −0.6
9 18 2.604 0.052 0.149 2.654 2.505
10 7 2.670 0.040 0.116 2.704 2.588
12 1 2.671 0.000 2.671 2.671
Bi5+ 4 1 1.979 0.000 1.979 1.979
6 10 2.110 0.013 0.038 2.130 2.092
[Figure 7]
Figure 7
Mean bond-length distributions for selected configurations of the metalloid ions bonded to O2−: (a) [3]B3+, (b) [4]B3+, (c) [4]Si4+, (d) [4]Ge4+, (e) [6]Ge4+, (f) [4]As5+, (g) [6]Sb5+, (h) [6]Te4+, (i) [7]Te4+, (j) [8]Te4+, (k) [6]Te6+.
[Figure 8]
Figure 8
Mean bond-length distributions for selected configurations of the post-transition metal ions bonded to O2−: (a) [4]Al3+, (b) [5]Al3+, (c) [6]Al3+, (d) [4]Ga3+, (e) [6]Ga3+, (f) [6]In3+, (g) [6]Sn4+, (h) [5]Pb2+, (i) [6]Pb2+, (j) [7]Pb2+, (k) [8]Pb2+, (l) [9]Pb2+, (m) [6]Bi3+, (n) [7]Bi3+, (o) [8]Bi3+.

Similar to the case for non-metal cations, both cations that form strongly bonded oxyanions (e.g. B3+, Si4+, As5+) and cations with `fully stereoactive' lone-pair electrons (i.e. with coordination numbers [3] and [4]) have a narrow range of mean bond lengths, typically ∼0.06–0.09 Å. This range is larger for ion configurations with stereoactive lone-pair electrons where secondary bonds are formed (e.g. [5–12]Bi3+), and is typically ∼0.1–0.3 Å for ion configurations with a dataset larger than ∼10 coordination polyhedra (mean bond-length range is highly dependent on sample size). The largest range of mean bond length observed is for [6]Te4+ with 0.573 Å, followed by [7]Te4+ with 0.380 Å and [8]Te4+ with 0.323 Å; however, this may due to the relatively high occurrence of ion configurations in coordination numbers ∼[6]–[8]. In comparison, the largest mean bond-length range observed for non-metal ions with stereoactive lone-pair electrons bonded to O2− is 0.227 Å for [6]Se4+ (Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]), and the largest range for the alkali and alkaline earth metals bonded to O2−, respectively, are 0.652 Å for [6]K+ and 0.436 Å for [10]Sr2+; the mean bond-length ranges are typically ∼0.3–0.4 Å for these two families (Gagné & Hawthorne, 2016a[Gagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602-625.]).

6.2.1. Bond-length distortion

We give the bond-length distortion plots for the metalloid and post-transition metal ions bonded to O2− in Figs. S7 and S8, and in Figs. 9[link] and 10[link] for those with adequate sample sizes. We use the definition of Brown & Shannon (1973[Brown, I. D. & Shannon, R. D. (1973). Acta Cryst. A29, 266-282.]) for distortion, i.e. the mean-square relative deviation of bond lengths from their average value. These plots show that mean bond length correlates highly with bond-length distortion for ion configurations observed with distortion values > 20 × 10−3, e.g. R2 = 0.92 for [6]Te4+, 0.86 for [8]Bi3+, but correlates poorly below that. A similar threshold was observed at ∼10 × 10−3 for the non-metal ions with stereoactive lone-pair electrons (Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]).

[Figure 9]
Figure 9
The effect of bond-length distortion on mean bond length for selected configurations of the metalloid ions bonded to O2−: (a) [3]B3+, (b) [4]B3+, (c) [4]Si4+, (d) [4]Ge4+, (e) [6]Ge4+, (f) [4]As5+, (g) [6]Sb5+, (h) [6]Te4+, (i) [7]Te4+, (j) [8]Te4+, (k) [6]Te6+.
[Figure 10]
Figure 10
The effect of bond-length distortion on mean bond length for selected configurations of the post-transition metal ions bonded to O2−: (a) [4]Al3+, (b) [5]Al3+, (c) [6]Al3+, (d) [4]Ga3+, (e) [6]Ga3+, (f) [6]In3+, (g) [6]Sn4+, (h) [5]Pb2+, (i) [6]Pb2+, (j) [7]Pb2+, (k) [8]Pb2+, (l) [9]Pb2+, (m) [6]Bi3+, (n) [7]Bi3+, (o) [8]Bi3+.
6.2.2. Factors affecting mean bond-length variations

A thorough investigation of potential factors leading to mean bond-length variation was done by Gagné & Hawthorne (2017[Gagné, O. C. & Hawthorne, F. C. (2017). Acta Cryst. B73, 1019-1031.]) for 55 ion configurations, which included analysis for [3]B3+, [4]B3+, [4]Al3+, [6]Al3+, [4]Si4+, [4]Ga3+, [4]Ge4+, [4]As5+, [6]Sb5+ and [6]Te6+. However, ion configurations with lone-pair electrons were not analyzed due to inadequate sample size. One of the conclusions of the study of Gagné & Hawthorne (2017[Gagné, O. C. & Hawthorne, F. C. (2017). Acta Cryst. B73, 1019-1031.]) is that the well ingrained correlation between mean bond length and mean coordination number of the bonded anions, proposed in the late 1960s, in fact resulted from small sample size, and is not of general applicability to inorganic oxide and oxysalt structures. They also confirmed bond-length distortion as a causal factor of mean bond-length variation and quantified its effect, and found no statistically significant correlation between mean bond length and the mean electronegativity and mean ionization energy of the next-nearest neighbours.

Let us examine the results for the metalloid and post-transition metal ions: [3]B3+ (n = 237 coordination polyhedra), [4]B3+ (n = 148), [4]Al3+ (n = 49), [6]Al3+ (n = 58), [4]Si4+ (n = 335), [4]Ga3+ (n = 27), [4]Ge4+ (n = 64), [4]As5+ (n = 59), [6]Sb5+ (n = 19) and [6]Te6+ (n = 21). Student t-tests show that for (1) bond-length distortion, (2) mean coordination number of bonded anion, (3) mean electronegativity and (4) mean ionization energy of the next-nearest neighbours, there are 16 of 40 possible correlations that are significant at the 95% confidence level. For bond-length distortion, they are for (R2) [4]B3+(0.33), [6]Al3+ (0.23), [6]Sb5+ (0.45) and [6]Te6+ (0.28); for mean coordination number of bonded anion, [3]B3+ (0.10), [4]B3+ (0.05), [4]Al3+ (0.17), [6]Al3+ (0.15) and [4]Ga3+ (0.29); for mean electronegativity of the next-nearest neighbours, [4]Ga3+ (−0.17), [4]As5+ (0.02) and [6]Te6+ (0.10); for mean ionization energy of the next-nearest neighbours, [4]Si4+ (−0.08), [4]Ga3+ (0.33), [4]As5+ (−0.09) and [6]Te6+ (−0.04). A negative symbol before R2 indicates that the observed correlation with mean bond length is negative.

As discussed by Gagné & Hawthorne (2017[Gagné, O. C. & Hawthorne, F. C. (2017). Acta Cryst. B73, 1019-1031.]), values of R2 and p-values vary significantly as a function of sample size (R2 values sometimes greater than 0.2 for sample sizes smaller than 100 coordination polyhedra for these variables), and although results for sample sizes > 35 coordination polyhedra are generally indicative, analysis of ion configurations with less than ∼100 coordination polyhedra cannot be considered statistically reliable. In the above case, the mean R2 values for the four variables considered are (1) 0.32, (2) 0.15, (3) −0.02, and (4) 0.04. Based on (1) lack of statistical significance in most cases, (2) low R2 values for those cases that are statistically significant, (3) the reliability of the R2 values based on the sample size study of Gagné & Hawthorne (2017[Gagné, O. C. & Hawthorne, F. C. (2017). Acta Cryst. B73, 1019-1031.]), and (4) a lack of demonstrated causality between mean bond length and these variables, we assume that mean bond length shows little or no correlation with the mean coordination number of bonded anion, the mean electronegativity of the next-nearest neighbours and the mean ionization energy of the next-nearest neighbours for the metalloid and post-transition metal ions.

The case for bond-length distortion is more interesting. It is clear from Figs. 9[link] and 10[link] that mean bond length is highly correlated to bond-length distortion for ion configurations that generally occur as highly distorted (e.g. ions with stereoactive lone-pair electrons), but correlates poorly otherwise. In addition, bond-length distortion is the only of the four potential factors analyzed above that has been demonstrated to be causal, via the distortion theorem (e.g. Brown & Shannon, 1973[Brown, I. D. & Shannon, R. D. (1973). Acta Cryst. A29, 266-282.]; Allmann, 1975[Allmann, R. (1975). Monatsh. Chem. 106, 779-793.]; Brown, 1978[Brown, I. D. (1978). Chem. Soc. Rev. 7, 359-376.]; Urusov, 2003[Urusov, V. S. (2003). Z. Kristallogr. 218, 709-719.]). Because of this, we can confidently say that bond-length distortion has a non-negligible effect on mean bond length for some strongly bonded metalloid and post-transition metal ions, and is the main cause of mean bond-length variation for highly distorted configurations of these ions.

Altogether, mean bond-length correlates poorly with the listed factors for the sample studied, and it is clear that one or more other factors affect mean bond-length variation. Following a study of a priori bond lengths in a variety of structures containing [4]Al3+, [6]Al3+ and [12]Ba2+, Gagné & Hawthorne (2017[Gagné, O. C. & Hawthorne, F. C. (2017). Acta Cryst. B73, 1019-1031.]) showed that a priori bond lengths do not correlate to observed bond lengths across structure types, although they are known to correlate well within structure types (e.g. R2 > 0.99 for milarite; Gagné & Hawthorne, 2016b[Gagné, O. C. & Hawthorne, F. C. (2016b). Can. Mineral. 54, 1229-1247.]). Following this, Gagné & Hawthorne (2017[Gagné, O. C. & Hawthorne, F. C. (2017). Acta Cryst. B73, 1019-1031.]) proposed that the inability of crystal structures to attain their ideal (a priori) bond lengths within the constraints of space-group symmetry is the leading cause of mean bond-length variation in crystals. As we concluded in the previous article of this series for the oxyanions of non-metals (Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]), this phenomenon seems plausible in explaining the mean bond-length variations observed here, and should be investigated further.

7. Summary

(1) We have examined the bond-length distributions for 33 configurations of the metalloid ions bonded to O2− using 5279 coordination polyhedra and 21 761 bond distances, and for 56 configurations of the post-transition metal ions bonded to O2− using 1821 coordination polyhedra and 10 723 bond distances.

(2) We find that for the seven metalloid and post-transition elements with lone-pair electrons we observe bonded to O2−, the most common state between their n versus n+2 oxidation states is that of higher oxidation state for As and Sb, and lower oxidation state for Sn, Te, Tl, Pb and Bi.

(3) We find no correlation between bond-valence sum and coordination number for cations with stereoactive lone-pair electrons using the bond-valence parameters of Gagné & Hawthorne (2015[Gagné, O. C. & Hawthorne, F. C. (2015). Acta Cryst. B71, 562-578.]).

(4) We confirm the absence of a correlation between lone-pair stereoactivity and coordination number when including secondary bonds, whereby both intermediate states of lone-pair stereoactivity and inert lone pairs may be observed for any coordination number > [4] of a cation with lone-pair electrons.

(5) We observe variations in mean bond lengths of ∼0.06–0.09 Å for strongly bonded oxyanions of metalloid and post-transition metal ions, and ∼0.1–0.3 Å for these ions that display lone-pair stereoactivity.

(6) We show that bond-length distortion is a leading cause of mean bond-length variation for ions with stereoactive lone-pair electrons, and that the causes of mean bond-length variation for strongly bonded cations (i.e. oxyanions) remain unclear. The most probable cause of mean bond-length variation for these ions is the effect of structure type, i.e. stress produced by the inability of a structure to follow its a priori bond lengths.

Supporting information


Funding information

This work was funded by an NSERC PGS-D3 Scholarship to OCG and UM Duff Roblin and GETS Fellowship from the University of Manitoba, and a Discovery grant to FCH from the Natural Sciences and Engineering Research Council of Canada.

References

First citationAdair, B. A., Diaz de Delgado, G., Delgado, J. M. & Cheetham, A. K. (2000). J. Solid State Chem. 151, 21–24.  Web of Science CrossRef CAS Google Scholar
First citationAlcock, N. W. (1972). Adv. Inorg. Radiat. Chem. 15, 1–57.  CAS Google Scholar
First citationAlig, H. & Trömel, M. (1992). Z. Kristallogr. 201, 213–222.  CrossRef CAS Web of Science Google Scholar
First citationAlimpiev, A. I., Merkulov, A. A., Solntsev, V. P., Tsvetkov, E. G., Matrosov, V. N. & Pestryakov, E. V. (2002). J. Cryst. Growth, 237, 884–889.  Web of Science CrossRef Google Scholar
First citationAllmann, R. (1975). Monatsh. Chem. 106, 779–793.  CrossRef CAS Web of Science Google Scholar
First citationBarbier, J. & Levy, D. (1997). Z. Kristallogr. 212, 519–528.  CrossRef CAS Web of Science Google Scholar
First citationBaur, W. H. (1978). Acta Cryst. B34, 1751–1756.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationBersuker, I. B. (1984). The Jahn–Teller Effect and Vibronic Interactions in Modern Chemistry. New York: Plenum.  Google Scholar
First citationBoiocchi, M., Callegari, C. & Ottolini, L. (2006). Am. Mineral. 91, 1170–1177.  Web of Science CrossRef CAS Google Scholar
First citationBrown, I. D. (1978). Chem. Soc. Rev. 7, 359–376.  CrossRef CAS Web of Science Google Scholar
First citationBrown, I. D. (1988). Acta Cryst. B44, 545–553.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationBrown, I. D. (2011). J. Phys. Chem. A, 115, 12638–12645.  Web of Science CrossRef CAS PubMed Google Scholar
First citationBrown, I. D. & Faggiani, R. (1980). Acta Cryst. B36, 1802–1806.  CSD CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationBrown, I. D. & Shannon, R. D. (1973). Acta Cryst. A29, 266–282.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationCalvo, C. & Faggiani, R. (1974). J. Chem. Soc. Chem. Commun. pp. 714–715.  CrossRef Web of Science Google Scholar
First citationCalvo, C., Faggiani, R. & Krishnamachari, N. (1975). Acta Cryst. B31, 188–192.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationChristy, A. G., Mills, S. J. & Kampf, A. R. (2016). Mineral. Mag. 80, 415–545.  Web of Science CrossRef CAS Google Scholar
First citationEmme, H. & Huppertz, H. (2003). Chem. Eur. J. 9, 3623–3633.  Web of Science CrossRef PubMed CAS Google Scholar
First citationFaggiani, R. & Calvo, C. (1976). Can. J. Chem. 54, 3319–3324.  CrossRef CAS Web of Science Google Scholar
First citationFleet, M. E. & Liu, X. (2003). Z. Kristallogr. 218, 795–801.  Web of Science CrossRef CAS Google Scholar
First citationGagné, O. C. (2018). Acta Cryst. B74, 49–62.  CrossRef IUCr Journals Google Scholar
First citationGagné, O. C. & Hawthorne, F. C. (2015). Acta Cryst. B71, 562–578.  Web of Science CrossRef IUCr Journals Google Scholar
First citationGagné, O. C. & Hawthorne, F. C. (2016a). Acta Cryst. B72, 602–625.  Web of Science CrossRef IUCr Journals Google Scholar
First citationGagné, O. C. & Hawthorne, F. C. (2016b). Can. Mineral. 54, 1229–1247.  Google Scholar
First citationGagné, O. C. & Hawthorne, F. C. (2017). Acta Cryst. B73, 1019–1031.  Web of Science CrossRef IUCr Journals Google Scholar
First citationGagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79–96.  CrossRef IUCr Journals Google Scholar
First citationGaly, J., Meunier, G., Andersson, S. & Åström, A. (1975). J. Solid State Chem. 13, 142–159.  CrossRef CAS Web of Science Google Scholar
First citationGillespie, R. J. (1972). Molecular Geometry. London: Van Nostrand Reinhold.  Google Scholar
First citationGillespie, R. J. & Nyholm, R. S. (1957). Quart. Rev. Chem. Soc. 11, 339–380.  CrossRef CAS Web of Science Google Scholar
First citationGuesdon, A., Monnin, Y. & Raveau, B. (2003). J. Solid State Chem. 172, 237–242.  Web of Science CrossRef CAS Google Scholar
First citationHawthorne, F. C., Burns, P. C. & Grice, J. D. (1996). Rev. Mineral. Geochem. 33, 41–115.  CAS Google Scholar
First citationHughes, K. A., Burns, P. C. & Kolitsch, U. (2003). Can. Mineral. 41, 677–685.  Web of Science CrossRef CAS Google Scholar
First citationJacobsen, S. D., Smyth, J. R. & Swope, R. J. (2003). Phys. Chem. Mineral. 30, 321–329.  Web of Science CrossRef CAS Google Scholar
First citationJansen, M. (1978). Angew. Chem. 90, 141–141.  CrossRef CAS Google Scholar
First citationJordan, T. H., Dickens, B., Schroeder, L. W. & Brown, W. E. (1980). Inorg. Chem. 19, 2551–2556.  CrossRef CAS Web of Science Google Scholar
First citationKahlenberg, V. (2001). Z. Anorg. Allg. Chem. 627, 2386–2390.  CrossRef CAS Google Scholar
First citationKahlenberg, V., Fischer, R. X. & Parise, J. B. (2000). J. Solid State Chem. 154, 612–618.  Web of Science CrossRef CAS Google Scholar
First citationKahlenberg, V., Lazic, B. & Krivovichev, S. V. (2005). J. Solid State Chem. 178, 1429–1439.  Web of Science CrossRef CAS Google Scholar
First citationKato, K., Takayama, E. & Kimizuka, N. (1979). Naturwissenschaften, 66, 616–617.  CrossRef CAS Web of Science Google Scholar
First citationLefebvre, I., Lannoo, M., Allan, G., Ibanez, A., Fourcade, J. & Jumas, J. C. (1987). Phys. Rev. Lett. 59, 2471–2474.  CrossRef PubMed CAS Web of Science Google Scholar
First citationLefebvre, I., Szymanski, M. A., Olivier-Fourcade, J. & Jumas, J. C. (1998). Phys. Rev. B, 58, 1896–1906.  Web of Science CrossRef CAS Google Scholar
First citationMagnusson, A. & Johansson, L. G. (1982). Acta Chem. Scand. A36, 429–433.  CrossRef CAS Web of Science Google Scholar
First citationMajzlan, J., Drahota, P. & Filippi, M. (2014). Rev. Mineral. Geochem. 79, 17–184.  Web of Science CrossRef Google Scholar
First citationMeier, S. F. & Schleid, T. (2006). Z. Anorg. Allg. Chem. 632, 2151–2151.  Google Scholar
First citationMüller-Buschbaum, H. & Szillat, H. (1994). Z. Anorg. Allg. Chem. 620, 642–646.  Google Scholar
First citationNarasaiah, T. V., Choudhary, R. N. P., Nigam, G. D. & Mattern, G. (1987). Acta Cryst. C43, 388–390.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationNikelski, T. & Schleid, T. (2003). Z. Anorg. Allg. Chem. 629, 2028–2033.  Web of Science CrossRef CAS Google Scholar
First citationOrgel, L. E. (1959). J. Chem. Soc. pp. 3815–3819.  CrossRef Web of Science Google Scholar
First citationPacalo, R. E. G. & Parise, J. B. (1992). Am. Mineral. 77, 681–684.  CAS Google Scholar
First citationPalkina, K. & Jost, K.-H. (1975). Acta Cryst. B31, 2281–2285.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationPico, C., Castro, A., Veiga, M. L., Gutierrez Puebla, E., Monge, M. A. & Ruiz Valero, C. (1986). J. Solid State Chem. 63, 172–178.  CrossRef CAS Web of Science Google Scholar
First citationPiffard, Y., Marchand, R. & Tournoux, M. (1975). Rev. Chim. Minér. 12, 210–217.  CAS Google Scholar
First citationPlatte, C. & Trömel, M. (1981). Acta Cryst. B37, 1276–1278.  CrossRef CAS IUCr Journals Google Scholar
First citationPloetz, K. B. & Müller-Buschbaum, H. (1982). Z. Anorg. Allg. Chem. 488, 38–44.  CAS Google Scholar
First citationPreiser, C., Lösel, J., Brown, I. D., Kunz, M. & Skowron, A. (1999). Acta Cryst. B55, 698–711.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationRichter, P. W., Kruger, G. J. & Pistorius, C. W. F. T. (1976). Acta Cryst. B32, 928–929.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationSchläger, M. & Hoppe, R. (1994). Z. Anorg. Allg. Chem. 620, 1867–1871.  Google Scholar
First citationSeshadri, R. & Hill, N. A. (2001). Chem. Mater. 13, 2892–2899.  Web of Science CrossRef CAS Google Scholar
First citationStearns, L. A., Groy, T. L. & Leinenweber, K. (2005). J. Solid State Chem. 178, 2594–2601.  Web of Science CrossRef CAS Google Scholar
First citationStoltzfus, M. W., Woodward, P. M., Seshadri, R., Klepeis, J.-H. & Bursten, B. (2007). Inorg. Chem. 46, 3839–3850.  Web of Science CrossRef PubMed CAS Google Scholar
First citationThauern, H. & Glaum, R. (2003). Z. Anorg. Allg. Chem. 629, 479–486.  Web of Science CrossRef CAS Google Scholar
First citationUrusov, V. S. (2003). Z. Kristallogr. 218, 709–719.  Web of Science CrossRef CAS Google Scholar
First citationWaghmare, U. V., Spaldin, N. A., Kandpal, H. C. & Seshadri, R. (2003). Phys. Rev. B, 67, 125111.  Web of Science CrossRef Google Scholar
First citationWalsh, A., Payne, D. J., Egdell, R. G. & Watson, G. W. (2011). Chem. Soc. Rev. 40, 4455–4463.  Web of Science CrossRef CAS PubMed Google Scholar
First citationWatson, G. W. & Parker, S. C. (1999). J. Phys. Chem. B, 103, 1258–1262.  Web of Science CrossRef CAS Google Scholar
First citationWatson, G. W., Parker, S. C. & Kresse, G. (1999). Phys. Rev. B, 59, 8481–8486.  Web of Science CrossRef 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 logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206
Follow Acta Cryst. B
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds