The trans effect: methodological musings

1992 ◽  
Vol 70 (10) ◽  
pp. 2574-2601 ◽  
Author(s):  
Osvald Knop ◽  
S. C. Choi ◽  
David C. Hamilton

The trans effect (TE) in the present context refers to the electronic compensation which in collinear homoligand L—Z—L* trans bond pairs lengthens the Z—L* bond when the Z—L bond is shortened. The existence of a functional relation d* = f(d) between the conjugated bond lengths d(Z—L) and d*(Z—L*) (d and d* not equivalent by symmetry; population A) has been demonstrated for a variety of Z-L combinations, with Z mostly from Groups VI and V and L mostly a halogen. The two model functions investigated in detail are the empirical DPF (difference power fit), d* – d0 = K(d − d0)−c, and the semiempirical CSBO (constant sum of bond orders) based on a modified 3-centre 4-electron bond concept, d* − d0 = −B ln {1 − expt[−(d − d0)/B]}, where B = b0 + b1(d − d0). Fitting DPF and CSBO to experimental d,d* data sets involves 3-parameter nonlinear optimization; in this CSBO differs from the 2-parameter treatment of Sheldrick etal., in which the limiting bond length d0 was supplied externally. Modified versions of DPF and CSBO have been devised to accommodate, along with A, d,d* pairs in which d = d* by symmetry (population S).The relative merits of DPF and CSBO and the various aspects of TE quantification are discussed at length, among these the effect of the oxidation state of Z and of the presence of heteroligands on Z. The meaning of the parameters of optimization and the existence of "chemical" trends between them are examined as well as the importance of the symmetrically balanced bond length de = d = d* and of the total d range Δ = de−d0 resulting from the d,d* regressions. Attempts to extend TE quantification to collinear heteroligand L1—Z—L2trans bond pairs have provided insight into the nature of the bond-length variation in such systems. The very good DPF and CSBO fits to d,d* sets obtained from 6-31G* optimizations of the equilibrium geometries of the OBOX, XOCN, and OCNY (X, Y = H, F, Cl, Li, Na, or no ligand) molecules and ions support the validity of the modified 3c4e model in accounting for the TE bond-length relationships.

2017 ◽  
Vol 259 ◽  
pp. 40-44 ◽  
Author(s):  
S.D. Singh ◽  
A.K. Poswal ◽  
C. Kamal ◽  
Parasmani Rajput ◽  
Aparna Chakrabarti ◽  
...  

2009 ◽  
Vol 24 (S10) ◽  
pp. 231-249
Author(s):  
M. J. Scanlan ◽  
I. H. Hillier ◽  
E. E. Hodgkin ◽  
R. P. Sidebotham ◽  
C. M. Warwick ◽  
...  
Keyword(s):  

2020 ◽  
Author(s):  
Olivier Charles Gagné ◽  
Frank Christopher Hawthorne

Bond-length distributions are examined for 63 transition-metal ions bonded to O2- in 147 configurations, for 7522 coordination polyhedra and 41,488 bond distances, providing baseline statistical knowledge of bond lengths for transi-tion metals bonded to O2-. A priori bond valences are calculated for 140 crystal structures containing 266 coordination poly-hedra for 85 transition-metal ion configurations with anomalous bond-length distributions. Two new indices, Δ𝑡𝑜𝑝𝑜𝑙 and Δ𝑐𝑟𝑦𝑠𝑡, are proposed to quantify bond-length variation arising from bond-topological and crystallographic effects in extended solids. Bond-topological mechanisms of bond-length variation are [1] non-local bond-topological asymmetry, and [2] multi-ple-bond formation; crystallographic mechanisms are [3] electronic effects (with inherent focus on coupled electronic-vibra-tional degeneracy in this work), and [4] crystal-structure effects. The Δ𝑡𝑜𝑝𝑜𝑙 and Δ𝑐𝑟𝑦𝑠𝑡 indices allow one to determine the primary cause(s) of bond-length variation for individual coordination polyhedra and ion configurations, quantify the dis-torting power of cations via electronic effects (by subtracting the bond-topological contribution to bond-length variation), set expectation limits regarding the extent to which functional properties linked to bond-length variations may be optimized in a given crystal structure (and inform how optimization may be achieved), and more. We find the observation of multiple bonds to be primarily driven by the bond-topological requirements of crystal structures in solids. However, we sometimes observe multiple bonds to form as a result of electronic effects (e.g. the pseudo Jahn-Teller effect); resolution of the origins of multiple-bond formation follows calculation of the Δ𝑡𝑜𝑝𝑜𝑙 and Δ𝑐𝑟𝑦𝑠𝑡 indices on a structure-by-structure basis. Non-local bond-topological asymmetry is the most common cause of bond-length variation in transition-metal oxides and oxysalts, followed closely by the pseudo Jahn-Teller effect (PJTE). Non-local bond-topological asymmetry is further suggested to be the most widespread cause of bond-length variation in the solid state, with no a priori limitations with regard to ion identity. Overall, bond-length variations resulting from the PJTE are slightly larger than those resulting from non-local bond-topological asym-metry, comparable to those resulting from the strong JTE, and less than those induced by π-bond formation. From a compar-ison of a priori and observed bond valences for ~150 coordination polyhedra in which the strong JTE or the PJTE is the main reason underlying bond-length variation, the Jahn-Teller effect is found not to have a symbiotic relation with the bond-topo-logical requirements of crystal structures. The magnitude of bond-length variations caused by the PJTE decreases in the fol-lowing order for octahedrally coordinated d0 transition metals oxyanions: Os8+ > Mo6+ > W6+ >> V5+ > Nb5+ > Ti4+ > Ta5+ > Hf4+ > Zr4+ > Re7+ >> Y3+ > Sc3+. Such ranking varies by coordination number; for [4], it is Re7+ > Ti4+ > V5+ > W6+ > Mo6+ > Cr6+ > Os8+ >> Mn7+; for [5], it is Os8+ > Re7+ > Mo6+ > Ti4+ > W6+ > V5+ > Nb5+. We conclude that non-octahedral coordinations of d0 ion configurations are likely to occur with bond-length variations that are similar in magnitude to their octahedral counterparts. However, smaller bond-length variations are expected from the PJTE for non-d0 transition-metal oxyanions.<br>


ACS Nano ◽  
2017 ◽  
Vol 11 (1) ◽  
pp. 501-515 ◽  
Author(s):  
Hendrik Naatz ◽  
Sijie Lin ◽  
Ruibin Li ◽  
Wen Jiang ◽  
Zhaoxia Ji ◽  
...  

1999 ◽  
Vol 86 (5) ◽  
pp. 2533-2539 ◽  
Author(s):  
M. Tormen ◽  
D. De Salvador ◽  
M. Natali ◽  
A. Drigo ◽  
F. Romanato ◽  
...  

1989 ◽  
Vol 162 ◽  
Author(s):  
Mark R. Pederson ◽  
Koblar A. Jackson ◽  
Warren E. Pickett

ABSTRACTIn order to gain insight into diamond growth, we have calculated equilibrium geometries for several adsorbates on the hydrogenated diamond <111> surface. While the adsorption height of a single methane radical onto a dangling bond is found to be in excellent agreement with the bulk-diamond bond length, the back bonded hydrogens of adjacent adsorbed methyl radicals repel one another. In contrast, adjacent acetlyinic radicals do not repel one another but lead to the introduction of double carbon bonds, misplaced carbon atoms above the active layer and a bond length which is too short in comparison to that of bulk diamond. Our calculations on the acetylene molecule near a dangling bond indicate that the resulting adsorbate bond length is substantially too large and that the carbon atom is unlikely to be stable directly above the surface carbon atom. Of the adsorbates studied, geometrical arguments suggest that the methyl radical is likely to be the most ideal adsorbate.


2014 ◽  
Vol 2 (14) ◽  
pp. 2475-2481 ◽  
Author(s):  
M. H. Du

Calculated Mn4+ emission energies for various oxides as functions of Mn–O bond length. The experimental values are shown (in red) wherever available. There are three groups of materials: the ones with small O–Mn–O bond angle distortion (black squares), the ones with large O–Mn–O bond angle distortion (blue circles), and phosphates (green triangles). Weak Mn4+-ligand hybridization as a result of long Mn–O bond lengths and/or large O–Mn–O bond angle distortion generally leads to higher emission energies.


Sign in / Sign up

Export Citation Format

Share Document