Skip to main content

Ligand Field Theory and the Fascination of Colours: Oxidic Iron(III) Solids as the Omnipresent Examples

  • Chapter
  • First Online:
50 Years of Structure and Bonding – The Anniversary Volume

Part of the book series: Structure and Bonding ((STRUCTURE,volume 172))

  • 2056 Accesses

Abstract

The treatment of the high-spin d5-configurated iron(III) cation in 6- and 4-coordinate ligand fields is a highly complex matter. The 6A1-ground state allows only spin-forbidden transitions, here of relevance to ten spin-reduced 4A1-, 4A2-, 4E(2x)-, 4T1(3x)-, 4T2(3x)-states, which are not easy to handle. Though the literature offers a series of carefully prepared solids with beautifully resolved ligand field spectra, the philosophy of utilising these in terms of their binding character, particularly in respect to the d-electron cloud density between cation and the anions, diverges. Accordingly, the magnitudes of reported ligand field parameters Δ and of the Racah parameters of interelectronic repulsion B and C, which parameterise the mentioned effects, differ, and comparisons become difficult. In this review we propose a well-founded and comprehensible calculational procedure, in order to, as the main matter of concern, convince the readers that the ligand field spectra also sensibly reflect finer perturbational details of local or even cooperative binding quality. The origin of the latter effects is from the chemical environment beyond the first coordination sphere of a central, say FeIIIL n -complex (n = 6,4) in an extended solid. Already subtle disturbances of this kind will modify the shade of colour. An essential point of the discussion is the symmetry analysis, which provides rigorous constraints and sets strict conditions on what can be experimentally observed. We will first discuss manganese(II) in oxidic solids. Because a divalent cation is associated with rather ionic binding properties towards oxygen, crystal field theory is the appropriate analytical instrument. Iron(III) provides a situation, which requires more sophistication and a refinement of the theory by taking also bond covalency into account. A symmetry-based reinterpretation of the additional absorptions in the d–d-spectra of FeIII and CrIII in corundum-type solids is presented. This treatment sheds new light on the finer roots of the impressive red-to-green colour change of Cr3+ in mixed crystals Al2−x Cr x O3 with increasing x. Particular examples are discussed, where absorptions due to octahedral and tetrahedral iron(III) overlap in the ligand field spectra of spinel- and garnet-type solids, which model the hue in a predictable way. Finally, though not directly related to the primary topic, the charge-transfer properties of oxidic iron(III) are briefly examined. These absorptions often stray far into the visible region, with a very frequently significant influence on the apparent hue.

Dirk Reinen is the only surviving member of the founding editorial board of Structure and Bonding in 1966. C.K. Jorgensen, J.B. Neilands, R.S. Nyholm and R.J.P. Williams sadly are no longer with us.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 169.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 219.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 219.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. Ilse FE, Hartmann H (1951) Z Phys Chem 197:239

    CAS  Google Scholar 

  2. Griffith JS (1971) The theory of transition metal ions. Cambridge University Press, London

    Google Scholar 

  3. Figgis BN, Hitchman MA (2000) Ligand field theory and its applications. Wiley-VCH, New York

    Google Scholar 

  4. Lever ABP (1984) Inorganic electronic spectroscopy. Elsevier, Amsterdam

    Google Scholar 

  5. Jorgensen CK (1966) Struct Bond 1:1

    Google Scholar 

  6. Jorgensen CK (1968) Oxidation numbers and oxidation states. Springer, New York

    Google Scholar 

  7. Reinen D, Schwab G, Günzler V (1984) Z Anorg Allg Chem 516:140

    Article  CAS  Google Scholar 

  8. Stout JW (1959) J Chem Phys 31:709

    Article  CAS  Google Scholar 

  9. Srivastava JP, Mehra A (1972) J Chem Phys 57:1587

    Article  CAS  Google Scholar 

  10. Allen GC, Warren KD (1971) Struct Bond 9:49

    Article  CAS  Google Scholar 

  11. Reinen D, Atanasov M, Köhler P, Babel D (2010) Coord Chem Rev 254:2703

    Article  CAS  Google Scholar 

  12. Reinen D (2014) Z Anorg Allg Chem 640:2677 (The author apologises for misprints in Fig 2: b 4Eg, b 4T1g should read as b 4T2g, b 4Eg, respectively)

    Google Scholar 

  13. Lehmann G (1970) Z Phys Chem 72:279

    Article  CAS  Google Scholar 

  14. Babel D (1967) Struct Bond 3:1

    Article  CAS  Google Scholar 

  15. Neuenschwander K, Güdel HU, Collingwood JC, Schatz PN (1983) Inorg Chem 22:1712

    Article  CAS  Google Scholar 

  16. Reinen D, Atanasov M, Lee SL (1998) Coord Chem Rev 175:91

    Article  CAS  Google Scholar 

  17. Atanasov M, Reinen D (2003) Comprehensive coordination chemistry, vol I fundamental. Elsevier, London, p 669

    Book  Google Scholar 

  18. Lenglet M, Bizi M, Jorgensen CK (1990) J Solid State Chem 86:82

    Article  CAS  Google Scholar 

  19. Weiss A, Witte H (1983) Kristallstruktur und Chemische Bindung. Verlag Chemie, Weinheim

    Google Scholar 

  20. Schmitz-DuMont O, Reinen D (1960) Z Elektrochem Ber Bunsenges Phys Chem 64:330

    CAS  Google Scholar 

  21. Reinen D (1969) Struct Bond 6:30

    Article  CAS  Google Scholar 

  22. Köhler P, Amthauer G (1979) J Solid State Chem 28:329

    Article  Google Scholar 

  23. Amthauer G, Günzler V, Hafner SS, Reinen D (1982) Z Kristallogr 161:167

    Article  CAS  Google Scholar 

  24. Moss SC, Newnham RE (1964) Z Kristallogr 120:359

    Article  CAS  Google Scholar 

  25. Pailhé N, Wattiaux A, Gaudon M, Demourgues A (2010) J Solid State Chem. doi:10.1016/j.jssc.2010.04.043

    Google Scholar 

  26. Lehmann G, Harder H (1970) Am Mineral 55:98

    CAS  Google Scholar 

  27. Ferguson J, Fielding PE (1972) Aust J Chem 25:1371

    Article  CAS  Google Scholar 

  28. Krebs JJ, Maisch WG (1971) Phys Rev B 4:757

    Article  Google Scholar 

  29. Duffy JA (1990) Bonding, energy levels and bands in inorganic solids. Longman, London

    Google Scholar 

  30. Huda MN, Walsh A, Yanta Y, Wei SH, Al-Jassim MM (2010) J Appl Phys 107:123712

    Article  Google Scholar 

  31. Nassau K (1983) The Physics and Chemistry of Color. Wiley (see particularly part III)

    Google Scholar 

  32. Moore CE (1952) Atomic energy levels, Vol II Nat Bur Stand Circ 467

    Google Scholar 

  33. Shannon RD, Prewitt CT (1969) Acta Crystallogr B 25:925

    Article  CAS  Google Scholar 

  34. Shannon RD (1976) Acta Crystallogr A 32:751

    Article  Google Scholar 

Download references

Acknowledgement

The technical assistance by Mrs. D. Kloss and Mrs. M. Köhler in transforming the manuscript and the figures, respectively, into a publishable shape is gratefully recognised.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dirk Reinen .

Editor information

Editors and Affiliations

Appendices

Appendix 1: The Crystal Field Treatment of a d5-Cation in a Cubic Environment

The state energies of a free d5-cation, in relation to its high-spin 6S-ground state, are, in terms of the Racah parameters B 0 and C 0 which parameterize the interelectronic repulsion [2]:

$$ \begin{array}{l}\to {}^4\mathrm{G}\kern0.5em 10{B}_0 + 5\ {C}_0;\kern1.75em \to {}^4\mathrm{P}\kern0.5em 7\ {B}_0 + 7{C}_0;\\ {}\to {}^4\mathrm{D}\kern0.5em 17{B}_0 + 5{C}_0;\kern2.23em \to {}^4\mathrm{F}\kern0.5em 22\ {B}_0 + 7{C}_0\end{array} $$
(6)

These states transform and split, by symmetry, in an octahedral crystal field into the components: 4A1g, 4Eg, 4T1g, 4T2g; 4T1g; 4Eg, 4T2g; 4A2g, 4T1g, 4T2g – where 6A1g is the (high-spin) ground state. The energies are (see also Eq. 6) [2]:

$$ \begin{array}{l}{}^4{\mathrm{A}}_{1\mathrm{g}}\kern0.5em 10B+5C;{\kern2.6em }^4{\mathrm{E}}_{\mathrm{g}}\ 10B+5C\\ {}{}^4{\mathrm{A}}_{2\mathrm{g}}\kern0.5em 22B+7C;\kern2.5em 17B+5C\end{array} $$
(7)

\( {}^4{\mathrm{T}}_{1\mathrm{g}}\left|\begin{array}{ccc}\hfill -\Delta -E\hfill & \hfill 3B\sqrt{2}\hfill & \hfill -C\hfill \\ {}\hfill 3B\sqrt{2}\hfill & \hfill 9B+C-E\hfill & \hfill -3B\sqrt{2}\hfill \\ {}\hfill -C\hfill & \hfill -3B\sqrt{2}\hfill & \hfill \Delta -E\hfill \end{array}\right|=0,\kern0.84em \mathrm{with}\ 10B + 6\ \mathrm{C}\ \mathrm{t}\mathrm{o}\ \mathrm{be}\ \mathrm{added}\ \mathrm{t}\mathrm{o}\ \mathrm{t}\mathrm{he}\ \mathrm{t}\mathrm{hree}\ \mathrm{solutions} \)

\( {}^4{\mathrm{T}}_{2\mathrm{g}}\left|\begin{array}{ccc}\hfill -\Delta -E\hfill & \hfill -\sqrt{6}B\hfill & \hfill -\left(4B+C\right)\hfill \\ {}\hfill -\sqrt{6}B\hfill & \hfill -5B-C-E\hfill & \hfill -\sqrt{6}B\hfill \\ {}\hfill -\left(4B+C\right)\hfill & \hfill -\sqrt{6}B\hfill & \hfill \varDelta -E\hfill \end{array}\right|=0,\kern0.85em \mathrm{with}\ \mathrm{diagonal}\ \mathrm{additions}\ \mathrm{of}\ 18B + 6C \)

Here, the ligand field parameter Δ and the modified Racah parameters B and C account for the energy modifications due to placing the cation into an external crystal field. Accordingly, ten S = 5/2 – to – S = 3/2-transitions may be observed at the most, from which one (4A2g) and those with a positive Δ dependence (c 4T1g, c 4T2g) are expected only at very high energies; furthermore, two (4A1g, a 4Eg) are degenerate in energy. The 4Tg-matrices indicate that – according to a (near-to) (−Δ)- or (+Δ)-dependence – the spin pairing may occur in the t2g-(t2g 4eg 1) or in the eg-subshell(t2g 2eg 3), respectively, while the b 4T1g- and b 4T2g-transitions are expected to depend only weakly on Δ. The specific symmetry of the two 4Tg matrices in respect to Δ is the reason why the Eq. (7) can be used in the case of tetrahedral crystal fields as well. Though usually a change of sign due to the reverse splitting into a lower-energy e- and a higher-energy t2-orbital set in Td as compared to Oh has to be applied, this is needless here.

Appendix 2: A Critical Review Towards the Numerical Instruments of Ligand Field Theory in Respect to Mn2+ and Fe3+ in Oh- and Td-Coordination

As has been discussed in detail elsewhere [7], spectral data for the free manganese(II)-cation [32], specifically the electronic transitions from 6S to 4G, 4P, 4D and 4F, can be rather precisely fitted to the Racah parameters B 0 and C 0, with numerical values of B 0 = 750 cm−1 and C 0/B 0 = 5.2 – though with the exception of the 6S → 4P-transition, which is calculated to occur at a, by about 10% too low energy. Thus – refraining from the doubtful-debatable procedure to introduce one additional parameter in order to achieve an improved fit to the experiment (Trees correction, see [7]) – one may confidently use the crystal field energies of Eq. (7) in Appendix 1 to parameterize experimental data, thereby leaving aside the 6A1g → 4T1g-transitions in the numerical calculation, however, for obvious reasons. The following master energy equations (utilising the B/C = B 0/C 0-ratio of 5.2 and referring to Δ/B-ratios in the range of those, deduced for oxidic manganese(II)-solids in Td(left) and Oh(right) crystal fields) are easily derived – restricting to the six lowest energy transitions [7]:

$$ {}^6{\mathrm{A}}_1{\to}^4{\mathrm{A}}_1,\;{\;_{\mathrm{a}}}^4\mathrm{E}\kern0.75em 36B\kern1.75em {{}_{\mathrm{b}}}^4\mathrm{E}\kern1em 43B $$
(8)

and, for Δ/B ≅ 7 (left) and ≅14 (right), respectively:

$$ \begin{array}{llll}{}^6{\mathrm{A}}_1\to \hfill & {{}_a}^4{\mathrm{T}}_1\hfill & -0.84\Delta +38.0B\hfill & -0.94\Delta +39.1B\hfill \\ {}\hfill & {{}_a}^4{\mathrm{T}}_2\hfill & -0.34\Delta +37.3B\hfill & -0.70\Delta +41.0B\hfill \\ {}\hfill & {{}_b}^4{\mathrm{T}}_2\hfill & -0.28\Delta +43.5B\hfill & -0.14\Delta +42.0B\hfill \\ {}\hfill & {{}_b}^4{\mathrm{T}}_1\hfill & +0.68\Delta +42.0B\hfill & +0.48\Delta +44.5B\hfill \end{array} $$

After all, one is in the comfortable situation, that there are frequently four experimental data (6A1 → 4A1, a 4E; b 4E; a 4T2; b 4T2), from which only two unknown parameters have to be deduced. The interelectronic repulsion parameter B can be derived from the Δ-independent transitions, where the minimum positions of the ground- and excited-state potential curves coincide. These minima are displaced in respect to each other, if Δ is involved. Though the Stoke shift, which is the energy difference between a transition in absorption and emission, is rather small in the case of a divalent cation such as Mn2+, the derived Δ parameter is nevertheless – though modestly – semi-empirical in nature. It is hence expected that – within the discussed limits – the nephelauxetic ratio β = B/B 0 images the covalency properties near to reality and that also Δ is a parameter, which reflects changes in the binding quality in a sensitive manner, if one sticks to the outlined critical calculation procedure. Having this in mind, both parameters, B and Δ, constitute as reliable members of the nephelauxetic and spectrochemical series of ligands [6].

Iron(III) differs from Mn(II) in its binding properties, by the distinctly enhanced covalent overlap between the metal and the ligands, and demands to classify the t2g- and eg-electrons in Oh according to their π- and σ-anti-bonding character, respectively; similarly, the e- and t2-electrons in Td are π-anti-bonding in the former and intermixed between σ- and π-anti-bonding in the latter MOs. Accordingly, the Racah parameters differ, depending on whether they refer to t2- or e-electrons. For example, in the case of Cr3+ in the octahedral coordination of fluoride elpasolites, B varies considerably between 790, ≅685 and ≅620 cm−1, for a t2g 3-, t2g 2eg 1- and t2g 1eg 2-configuration, respectively [11]. Having covalence phenomena of this type in mind, a coarser approach than the one introduced for Mn2+ seems appropriate – and this with the clear intention to not overload the treatment with too many not sufficiently meaningful parameters. We will accordingly follow an approach, in which only a singular B-value is used and in which also the 6A1 → 4T1-transitions are included into the fitting procedure, thereby utilising a C/B-ratio of 5.0. Though well defined, this approach certainly implies a larger uncertainty in the numerical fixation of the ligand field parameters. Nevertheless, “quod est demonstrandum”, rather satisfactory fits to the iron(III)-ligand field spectra are possible, with consistent interpretations of the derived binding parameters Δ and B. The basic demand is, as always, the assistance of an appropriate symmetry classification. As for Mn2+, we use in the following master equations for the transition energies of iron(III) in Oh- and Td-ligand fields at (for the Δ-dependent absorption bands) Δ/B-ratios, which are particularly applicable to the oxidic iron(III) solids, considered here, with C/B = 5.0:

$$ {}^6{\mathrm{A}}_1\to {}^4{\mathrm{A}}_1,\;{\;_{\mathrm{a}}}^4\mathrm{E}\ :\kern0.5em 35B\kern2.5em \to {\;_{\mathrm{b}}}^{\mathrm{a}}\mathrm{E}\ :\ 42B $$
(9)

and, for Δ/B ≅ 12 (left) and 20 (right), respectively:

$$ \begin{array}{llll}{}^6{\mathrm{A}}_1\to \hfill & {{}_{\mathrm{a}}}^4{\mathrm{T}}_1\hfill & -0.93\Delta +37.7B\hfill & -0.97\Delta +38.3B\hfill \\ {}\hfill & {{}_{\mathrm{a}}}^4{\mathrm{T}}_2\hfill & -0.77\Delta +41.7B\hfill & -0.90\Delta +43.7B\hfill \\ {}\hfill & {{}_{\mathrm{b}}}^4{\mathrm{T}}_2\hfill & -0.02\Delta +38.3B\hfill & -0.01\Delta +38.1B\hfill \\ {}\hfill & {{}_{\mathrm{b}}}^4{\mathrm{T}}_1\hfill & +0.57\Delta +41.8B\hfill & +0.23\Delta +47.2B\hfill \end{array} $$

Appendix 3: The Symmetry Relations and the d-Electron Density Distribution of Chromium(III) and Iron(III) in the Corundum Structure

If one marks the t2g- and eg-wave functions in Oh with ζ, η, ξ and Θ, ε, respectively, the following linear combinations result, if the axis of quantisation is now C3 instead of C4 (see column C3v (1)):

$$ \begin{array}{lll}{\mathrm{O}}_{\mathrm{h}}\hfill & {\mathrm{C}}_{3\mathrm{v}}(1)\hfill & {\mathrm{C}}_{3\mathrm{v}}(2)\hfill \\ {}{\mathrm{t}}_{2\mathrm{g}}\kern1.5em \zeta \hfill & 1/\sqrt{3}\left(\zeta +\eta +\xi \right)\kern1.5em {\mathrm{a}}_1\hfill & {{\mathrm{d}}_z}^2\hfill \\ {}\kern2.62em \eta \hfill & 1/\sqrt{6}\left(2\zeta -\eta -\xi \right)\kern1.12em \mathrm{e}\hfill & {\mathrm{d}}_{x^2-{y}^2}\hfill \\ {}\kern2.62em \xi \hfill & 1/\sqrt{2}\left(\eta -\xi \right)\hfill & {\mathrm{d}}_{xy}\hfill \\ {}{e}_g\kern1.5em \varTheta \hfill & \kern2em \varTheta \kern5.8em \mathrm{e}\hfill & {\mathrm{d}}_{xz}\hfill \\ {}\kern2.62em \varepsilon \hfill & \kern2em \varepsilon \kern6.1em \hfill & {\mathrm{d}}_{yz}\hfill \end{array} $$
(10)

Adopting the C3- as the new molecular z-axis, the d-wave functions transform as listed under C3v (2); the latter are the ones used in the main text and in Fig. 11. According to the presence of the threefold axis, any cyclic permutation of the wave functions under C3v (1) is also a valid choice. The d z 2-wave function models the metal pair interaction in the corundum lattice via a1. The two e-doublets come out to be non-bonding (from t2g) and σ-anti-bonding (from eg), respectively, if otherwise the octahedral symmetry is retained – as this is nearly so in Cr2O3. In difference, the structural C3v distortion is very significant in α-Fe2O3, and the two e-orbital sets are expected to intermix. The distance between the two, displaced, octahedral cations within the O3MO3MO3-binuclear pairs (Fig. 10) along the C3-axis is about 2.12, 2.17 and 2.27 Å for M being Al(III), Cr(III) and Fe(III), respectively. These values coarsely follow the sequence of the octahedral ionic radii, which are reported to be [33, 34] 0.535, 0.615 and 0.645 Å, only in the case of the two 3d-cations. For Al(III) this spacing is about 0.1 Å larger, which becomes apparent, if one corrects the internuclear distances in respect to the ionic radii differences. This observation indeed matches with the lack of intercationic binding in that case. Focussing on Cr(III) and Fe(III) accordingly, one further notes that the bond lengths towards the oxygen ligands in the boundary faces of the binuclear pairs (1.97 and 1.945 Å, respectively) indicate a by nearly 3% (again radii-corrected) reduced effective value in the latter case. Recalling the spectral results in addition, one has to conclude as follows: The distinct C3v- structural component, considering α-Fe2O3, has introduced, via the e(t2g) – e(eg)-interaction (Fig. 11, left), particularly a more pronounced t2g-splitting; this energy effect is nicely imaged via the by a factor of two enhanced parameter δ (see Eq. (3) and Table 3 in Sect. 3.3), and attended by the conversion of the two symmetry-equivalent orbital doublets into linear combinations (see Eqs (10)). Also the derived ligand field parameters (Δ, in 103 cm−1) for the here and in [21] discussed spinel-, garnet- and corundum-type phases indicate the exceptional position of α-Fe2O3:

$$ \begin{array}{lll}\mathrm{Compound}\hfill & \Delta \hfill & B\hfill \\ {}{\mathrm{ZnCr}}_2{\mathrm{O}}_4\hfill & 17.4\hfill & 0.64\hfill \\ {}{\mathrm{ZnFe}}_2{\mathrm{O}}_4\hfill & 15.7\hfill & 0.65\hfill \\ {}{\mathrm{Y}}_3{\mathrm{Cr}\mathrm{Ga}}_4{\mathrm{O}}_{12}\hfill & {16.1}_5\hfill & 0.63\hfill \\ {}{\mathrm{Y}}_3{\mathrm{Fe}\mathrm{Ga}}_4{\mathrm{O}}_{12}\hfill & 15.3\hfill & 0.65\hfill \\ {}\upalpha \hbox{-} {\mathrm{Cr}}_2{\mathrm{O}}_3\hfill & 15.8\hfill & 0.57\hfill \\ {}\upalpha \hbox{-} {\mathrm{Fe}}_2{\mathrm{O}}_3\hfill & =11.7\hfill & =0.60\hfill \end{array} $$
(11)

It is, induced by the dramatic rise of the intercationic overlap parameter δ, the pronounced decrease of Δ, which mirrors the binding landscape specifically in the α-Fe x Al2−x O3-mixed crystals (see Table 3). Δ becomes smaller by at least 20%, when replacing chromium(III) by iron(III) in the corundum structure, in comparison to 10% or even less in spinel- or garnet-type phases. Considering the above outlined model arguments, this is an expected experimental result. Already discussed is the reduction of B (in 103 cm−1, see Eq. 11) by about 10%, which is observed, if the within-pair overlap comes up in addition (Table 3). We think that – though the iron(III) bonding properties in the corundum lattice are highly complex – the proposed model is an acceptable symmetry-adjusted approach to what one may call the truth.

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Köhler, P., Reinen, D. (2016). Ligand Field Theory and the Fascination of Colours: Oxidic Iron(III) Solids as the Omnipresent Examples. In: Mingos, D. (eds) 50 Years of Structure and Bonding – The Anniversary Volume. Structure and Bonding, vol 172. Springer, Cham. https://doi.org/10.1007/430_2015_200

Download citation

Publish with us

Policies and ethics