Coordination geometry
Coordination geometry refers to the spatial arrangement of donor atoms from ligands around a central atom, usually a metal ion, in a coordination entity, defining the three-dimensional structure of the complex.[1] The coordination number, which is the number of such donor atoms bonded to the central atom via sigma bonds, primarily dictates the possible geometries adopted by the complex.[1] Common coordination numbers for transition metal complexes range from 2 to 9, with 4 and 6 being the most prevalent.[2] For coordination number 4, the typical geometries are tetrahedral and square planar, while for coordination number 6, the dominant arrangement is octahedral; other notable geometries include trigonal bipyramidal and square pyramidal for coordination number 5.[1] These preferences arise from a balance of electronic factors, such as the d-electron configuration of the metal, and steric effects from ligand size and shape, with octahedral geometry being especially common for first-row d-block metals due to its stability.[2] Coordination geometry plays a crucial role in determining the physical and chemical properties of coordination compounds, including their color, magnetic behavior, reactivity, and potential for stereoisomerism.[1] For instance, square planar complexes of d8 metals like Pt(II) often exhibit distinct reactivity patterns compared to tetrahedral d10 analogs, influencing applications in catalysis and materials science.[2] In nomenclature, geometries are specified using polyhedral symbols, such as OC-6 for octahedral, to precisely describe the structure and facilitate comparison across compounds.[1]Fundamentals
Definition and Scope
Coordination geometry refers to the three-dimensional spatial arrangement of ligands surrounding a central metal atom or ion in a coordination complex, typically forming polyhedral structures such as octahedra or tetrahedra.[3][4] This arrangement is primarily determined by the coordination number, which indicates the number of ligand donor atoms bound to the central atom, and influences the overall shape of the complex.[3] The scope of coordination geometry encompasses both mononuclear complexes, featuring a single central metal ion, and polynuclear complexes, which involve multiple metal centers linked by bridging ligands.[3] Unlike molecular geometry, which applies valence shell electron pair repulsion (VSEPR) theory to predict shapes based on electron pairs around an atom, coordination geometry emphasizes the positioning of metal-ligand bonds, often disregarding lone pairs on the metal.[3] This geometric framework plays a crucial role in the properties of coordination compounds, dictating their reactivity through factors like ligand accessibility and steric effects, as seen in isomer-specific reaction pathways.[3] It also governs stability, with chelating ligands forming more robust polyhedra due to the chelate effect, and impacts spectroscopic behavior by influencing d-orbital splitting, which in turn affects electronic transitions responsible for color.[3] Additionally, coordination geometry determines magnetic properties by shaping the electron configuration in the metal's d-subshell.[3] The foundational understanding of coordination geometry originated with Alfred Werner's early 20th-century studies on octahedral and tetrahedral arrangements in coordination complexes, for which he received the 1913 Nobel Prize in Chemistry.[5]Coordination Number
The coordination number (CN) in coordination chemistry refers to the number of donor atoms from ligands that form coordinate covalent bonds—typically sigma bonds—with the central metal atom.[6] For transition metal complexes, the CN typically ranges from 2 to 12, though values from 1 to 12 have been characterized.[6] Higher CNs are rare but possible in larger ions or specialized clusters, with the record at 16 observed in the cobaltaborane anion [CoB_{16}]^{-}.[7] The CN is influenced by several key factors. The ionic radius of the central metal ion is primary; larger radii permit higher CNs by providing more space around the metal for ligand attachment, following principles akin to those in ionic crystal packing.[8] Ligand size and type also play a critical role: small monodentate ligands favor higher CNs, while bulky ligands impose steric hindrance that reduces the CN to minimize repulsive interactions.[9] Polydentate ligands, which bind through multiple donor atoms, can effectively increase or stabilize the CN beyond what monodentate ligands allow.[10] Additionally, the electronic configuration of the metal ion affects the CN, as specific d-electron counts determine the availability of orbitals for bonding and the stability of resulting arrangements./05:Coordination_Chemistry_I-_Structures_and_Isomers/5.03:_Coordination_Numbers_and_Structures) Charge on the metal and ligands further modulates this, with higher metal charges often promoting higher CNs to achieve charge balance.[10] Representative examples illustrate these trends. Most d-block metals achieve a CN of 6 with small ligands, as in [Co(NH_3)_6]^{3+}, where the compact ammonia ligands fit around the cobalt(III) ion without significant steric issues.[11] A CN of 4 is common with similar small ligands but under conditions favoring lower coordination, such as in [ZnCl_4]^{2-}, influenced by the d^{10} electronic configuration of zinc(II).[12]Structural Features
Ligand Arrangement and Bond Angles
In coordination geometry, ligands are arranged around the central metal ion at the vertices of regular polyhedra to minimize electrostatic repulsions between them and maximize the overall symmetry of the complex. This arrangement follows principles analogous to the valence shell electron pair repulsion (VSEPR) theory, where the positions of ligands are determined by the need to space them as far apart as possible, thereby reducing ligand-ligand interactions and stabilizing the structure. The specific polyhedral geometry adopted depends on the coordination number, with ideal configurations achieving high symmetry that reflects the uniform distribution of ligands. For common geometries, the linear arrangement (coordination number 2) features two ligands positioned opposite each other, resulting in a bond angle of 180°, and belongs to the D∞h point group, characterized by infinite rotation axes and perpendicular mirror planes.[13] The trigonal planar geometry (coordination number 3) places three ligands in a plane with bond angles of 120°, exhibiting D3h symmetry, which includes a principal threefold rotation axis and horizontal mirror planes. In tetrahedral geometry (coordination number 4), four ligands occupy the vertices of a regular tetrahedron, with ideal bond angles given by \theta = \cos^{-1}\left(-\frac{1}{3}\right) \approx 109.47^\circ, derived from the vector sum of bond vectors meeting at the center such that their scalar product yields this value; this configuration has Td symmetry, featuring four threefold axes and mirror planes through edges.[13] Square planar geometry, another option for coordination number 4, arranges four ligands in a plane with adjacent bond angles of 90° and trans angles of 180°, belonging to the D4h point group with a fourfold rotation axis and multiple mirror planes. The octahedral geometry (coordination number 6) is one of the most prevalent, with six ligands at the vertices of a regular octahedron, featuring cis bond angles of 90° and trans angles of 180°, and Oh symmetry that includes threefold and fourfold rotation axes, inversion center, and extensive mirror planes. These ideal angles and symmetries provide the foundational framework for understanding coordination structures, influencing properties such as reactivity and spectroscopy.[13]Distortions and Variations
Coordination geometries in metal complexes often deviate from ideal polyhedral arrangements due to various factors, leading to distortions such as elongation, compression, and twisting of bond lengths and angles. Elongation typically involves lengthening of axial bonds in octahedral complexes, while compression shortens them, and twisting adjusts the rotational alignment of ligands relative to the metal center. These distortions stabilize the complex by reducing electronic or steric repulsion.[14] A prominent example is the Jahn-Teller distortion in octahedral complexes with degenerate electronic states, such as d⁹ Cu(II) systems where uneven occupancy of e_g orbitals (one with two electrons, the other with one) drives axial elongation to lower the energy of the doubly occupied orbital. This effect, predicted by the Jahn-Teller theorem, removes orbital degeneracy in nonlinear molecules, resulting in lower symmetry geometries like tetragonal distortion.[15][14] The energy stabilization from this distortion arises from vibronic coupling between electronic and vibrational modes; in the linear approximation for the E ⊗ e_g Jahn-Teller problem in octahedral fields, the stabilization energy is given by \Delta E = -\frac{3}{16} \frac{\lambda^2}{\Delta}, where λ is the linear electron-vibration coupling constant and Δ is the crystal field splitting. To derive this, consider the Hamiltonian including vibronic interaction: the undistorted energy is degenerate at E=0, but distortion along a normal mode Q introduces a linear term λ Q (σ_x), where σ_x is the Pauli matrix for the electronic doublet. Minimizing the total energy E(Q) = (1/2) k Q² - λ |Q| (for the lower sheet) yields Q_min = λ / k and ΔE = - λ² / (2k). For octahedral e_g modes, symmetry factors and the two-mode nature adjust the coefficient to 3/16 when expressing k in terms of Δ.[15] Distortions arise from electronic causes, such as uneven d-orbital filling leading to preferential occupation and bond weakening (e.g., in high-spin d⁴ Mn(III) complexes showing either elongation or compression along e_g axes). Steric causes involve large ligands that impose spatial constraints, promoting fluxional behavior where ligands rearrange to minimize repulsion, as seen in complexes with bulky phosphines distorting octahedral geometries toward square pyramidal forms. Relativistic effects in heavy metals, like gold and mercury, contract s-orbitals and enhance spin-orbit coupling, favoring linear coordination (e.g., two-coordinate Au(I)) over higher geometries due to stabilized 6s electrons.[16][17] Variations in geometry include dynamic fluxional processes, such as Berry pseudorotation in five-coordinate trigonal bipyramidal species (e.g., Fe(CO)₅), where axial and equatorial ligands interchange via a square pyramidal transition state with low barriers (2–7 kcal/mol), effectively averaging positions on NMR timescales. Chelate ring constraints further induce non-ideal angles; the bite angle (metal-donor-metal) of bidentate ligands, determined by backbone flexibility, deviates from 90° in octahedral sites (e.g., 82–85° for ethylenediamine), enforcing twists or flattenings in the polyhedron to accommodate ring strain.[18]Catalog of Geometries
Common Geometries for Low Coordination Numbers
For coordination number 2, the linear geometry is the predominant arrangement, characterized by a bond angle of 180° between the two ligands and the central metal ion. This configuration is common for d¹⁰ metal ions, such as Ag(I), where the filled d-shell minimizes electronic repulsion and favors collinear ligand placement. A representative example is the [Ag(NH₃)₂]⁺ complex, in which ammonia ligands coordinate to silver in a linear fashion, as confirmed by crystallographic studies.[19][20] The linear structure can be visualized as a simple diatomic-like arrangement, with the metal at the center and ligands at opposite ends of an axis, exhibiting D∞h point group symmetry. Coordination number 3 typically adopts either trigonal planar or T-shaped geometries, depending on the electronic configuration of the metal. In trigonal planar geometry, the three ligands lie in a single plane with bond angles of 120°, providing high symmetry (D₃h point group) and minimal steric strain for main-group or soft metals. The [HgI₃]⁻ anion exemplifies this, where mercury(II) is surrounded by three iodide ligands in a planar triangle, stabilized by the large size and polarizability of Hg(II).[21] This arrangement is often seen in complexes with d¹⁰ metals or those involving bulky halides. In contrast, T-shaped geometry arises as a distortion in d⁸ square-planar precursors, featuring bond angles of approximately 90° and 180°, with the axial ligands nearly linear and the equatorial one perpendicular. This is prevalent in 14-electron d⁸ transition metal complexes, such as certain Pt(II) species, where the electronic preference for low-spin configurations drives the deviation from planarity.[22] The T-shape can be depicted as a central metal with two ligands in a straight line and a third perpendicular, resembling a "T" in projection. For coordination number 4, the two primary geometries are tetrahedral and square planar, distinguished by their symmetry and electronic factors. Tetrahedral geometry (T_d point group) features bond angles of about 109.5°, with ligands at the vertices of a tetrahedron around the central metal, common for first-row transition metals with weak-field ligands that promote high-spin configurations. The [NiCl₄]²⁻ complex illustrates this, where Ni(II) (d⁸) adopts tetrahedral coordination due to chloride's weak ligand field, resulting in two unpaired electrons and paramagnetism. Visualized as a three-dimensional pyramid with the metal at the center, this structure avoids close ligand-ligand repulsions for smaller ions. Square planar geometry (D_{4h} point group), however, arranges four ligands in a plane with 90° bond angles, favored by d⁸ metals in the second and third rows where large crystal field splitting (Δ) exceeds pairing energy, stabilizing low-spin diamagnetic states. The [PtCl₄]²⁻ ion exemplifies this, with Pt(II) (d⁸) forming a flat square despite chloride ligands, owing to platinum's strong relativistic effects and extended d-orbitals enhancing splitting.[23] Factors favoring square planar over tetrahedral include stronger ligand fields, higher oxidation states, and 4d/5d metals, while tetrahedral is preferred for 3d metals with weak fields to minimize pairing costs.[24] The square planar form appears as a flat cross in diagrams, highlighting its planarity and potential for π-bonding interactions.Common Geometries for Higher Coordination Numbers
For coordination number 5 (CN=5), the two primary geometries are trigonal bipyramidal and square pyramidal, which differ in ligand positioning and symmetry. In trigonal bipyramidal geometry, three ligands occupy equatorial positions forming a plane with 120° angles, while two axial ligands sit above and below at 90° to the equator; this arrangement exhibits D_{3h} point group symmetry in ideal cases, as seen in iron pentacarbonyl, [Fe(CO)5], where the carbonyl ligands adopt this structure in the gas phase.[25] Square pyramidal geometry features four basal ligands in a square and one apical ligand, with C{4v} symmetry; a representative example is vanadyl acetylacetonate, [VO(acac)_2], where the vanadium(IV) center coordinates to two bidentate acetylacetonate ligands and an oxo group in a distorted square pyramidal arrangement.[26] These geometries arise from the electronic preferences of d-block metals, with trigonal bipyramidal being more common for dsp^3 hybridization in fluxional systems. Coordination number 6 (CN=6) predominantly adopts octahedral geometry, the most stable and ubiquitous for transition metal complexes due to its high symmetry and minimal ligand repulsion. In this arrangement, six ligands surround the central metal at the vertices of an octahedron, with ideal bond angles of 90° and Oh point group symmetry, exemplified by hexaamminecobalt(III), [Co(NH_3)6]^{3+}, where the cobalt(III) ion achieves a near-perfect octahedral coordination with ammonia ligands. For complexes with three identical bidentate or monodentate ligands, geometrical isomerism occurs: the facial (fac) isomer positions the three identical ligands on one triangular face (C{3v} symmetry), while the meridional (mer) isomer arranges them in a linear plane (C_{2v} symmetry); these isomers, such as in [Co(NH_3)_3(NO_2)_3], exhibit distinct spectroscopic and reactivity properties due to differences in ligand-metal interactions.[27] Higher coordination numbers (CN=7 and above) are less common in early transition metals but prevalent in larger ions like those of lanthanides and actinides, where increased ionic radii accommodate more ligands with geometries derived from polyhedral expansions. For CN=7, pentagonal bipyramidal geometry features five equatorial ligands in a pentagon and two axial positions (D_{5h} symmetry), as observed in the zirconate anion [ZrF_7]^{3-}, where the zirconium(IV) center bonds to seven fluorides in this configuration within ammonium or potassium salts. An alternative CN=7 geometry is the capped octahedron (C_{3v} symmetry), where a seventh ligand caps one face of an octahedral arrangement, often seen in early transition metal fluorides or carbonyls. For CN=8, common structures include the dodecahedron (D_{2d} symmetry) and square antiprism (D_{4d} symmetry), both minimizing steric strain; the octacyanomolybdate(IV) anion, [Mo(CN)_8]^{4-}, adopts a dodecahedral geometry in solid-state compounds, with molybdenum-carbon distances varying slightly to accommodate the cyanide ligands.[28] In lanthanides and actinides, coordination numbers often exceed 8 (up to 12), favoring tricapped trigonal prismatic or other high-symmetry polyhedra due to large cation sizes and ionic bonding dominance, contrasting with the lower coordination preferences in d-block metals.[29][30] For example, the [Ce(NO_3)_6]^{2-} anion exhibits icosahedral (I_h) geometry with coordination number 12.| Coordination Number | Common Geometry | Example | Point Group Symmetry |
|---|---|---|---|
| 5 | Trigonal bipyramidal | [Fe(CO)_5] | D_{3h} |
| 5 | Square pyramidal | [VO(acac)_2] | C_{4v} |
| 6 | Octahedral | [Co(NH_3)_6]^{3+} | O_h |
| 7 | Pentagonal bipyramidal | [ZrF_7]^{3-} | D_{5h} |
| 7 | Capped octahedral | Early metal fluorides (e.g., [NbF_7]^{2-}) | C_{3v} |
| 8 | Dodecahedral | [Mo(CN)_8]^{4-} | D_{2d} |
| 8+ | Icosahedral or higher polyhedra | Lanthanide/actinide complexes (e.g., [Ce(NO_3)_6]^{2-}) | Variable (e.g., I_h) |