Mössbauer Spectroscopy Rudolf L. Mössbauer 1929 ‐ 2011 1958  Recoilless Nuclear Resonance  Absorption of Gamma‐Radiation = the Mössbauer Effect (during PhD) 1961 Nobel Prize in Physics The Mössbauer resonance line is extremely narrow and allows hyperfine interactions to be resolved and evaluated, quadrupole splitting  and an isomer shift  1 The Mössbauer Effect 2 Recoilless Nuclear Resonance Absorption and Fluorescence of -Radiation 3 2 2 2mc E ER   ge EEE 0 REEE  0 Nuclear Decay Scheme for 57Fe Mössbauer Resonance 4 15 % Mössbauer Emission Radioactive 57Co (halflife 270 d) is generated in a cyclotron and  diffused into Rh serves as the gamma radiation  source for 57Fe Mössbauer spectroscopy 10 ns, 122 keV Nuclear Resonance Absorption 5 Mössbauer Active Elements 6 Nuclides suitable for Mössbauer spectroscopy: Excited nuclear state lifetimes 10‐6 ‐ 10‐11 s  Transition energies 5 ‐180 keV.  Longer (shorter) lifetimes = too narrow (broad) emission and absorption lines, no  effective overlap Transition energies > 180 keV = too large recoil, < 5 keV = absorbed in the source and  absorber material The Mössbauer effect detected in 80 isotopes of 50 elements,  only 20 elements studied in practice Fe, Sn, Sb, Te, I, Au, Ni, Ru, Ir, W, Kr, Xe, rare earth elements, Np. More than 90 % of publications refer to 57Fe  Nuclear Parameters for Selected Mössbauer Isotopes 7EC = electron capture, β–=beta-decay, IT = isomeric transition, α = alpha-decay Mössbauer Source Activity = 10 ‐270 mCi, lifetime 10 y 1 Ci = 37 GBq Reference Absorbers 8 Mössbauer Source 119mSn source = CaSnO3 matrix with 119mSn  Activity = 2 ‐ 40 mCi, lifetime 10 y 9 Mean Lifetime  of Excited State and Natural Line Width  10 An excited state (nuclear or electronic) of mean lifetime  does not have a sharp energy value, but only a value within the energy range E the Heisenberg Uncertainty Principle:  E  t  ħ natural line width  = ħ /  Transitions from an excited (e) to the ground state (g) involve all energies within the range of E The transition probability or intensity as a function of E = a spectral line centered around the most probable transition energy E0 Transition Probability 11 Lorentzian formula for spectral line shape Resonance absorption is observable only if the emission and absorption lines overlap sufficiently. This is not the case when the lines are too narrow or too broad Suitable for Mössbauer spectroscopy: lifetime = 10-6 - 10-11 s The width at half maximum of the spectral line = natural line width  determined by the mean lifetime  57Fe (14.4 keV)  =1.43 10-7 s  = 4.6 10-9 eV Recoil Effect 12 57Fe (14.4 keV)  = 4.6 10-9 eV Recoil energy ER = 2 10-3 eV Much larger (5‐6 orders of magnitude) than the natural line width = no resonance possible between free atoms The Mössbauer effect cannot be observed for freely moving atoms or molecules, i.e. in gaseous or liquid state 2 2 2mc E ER   Recoil Effect 13 57Fe (14.4 keV)  = 4.6 10-9 eV Recoil energy ER = 2  10-3 eV Much larger (5‐6 orders of magnitude) than the natural line width = no resonance possible between free atoms The Mössbauer effect cannot be observed for freely moving atoms or molecules, i.e. in gaseous or liquid state 2 2 2mc E ER   Recoilless Emission and Absorption 14 Solid state, crystalline or non‐crystalline = atoms tightly bound in the lattice Unshifted transition lines overlap = nuclear resonance absorption observed Large mass M of a solid particle as compared to an atom = the linear momentum created by emission and absorption of a gamma quantum practically vanishes The recoil energy is mostly transferred to the lattice vibrational system.  The vibrational energy of the lattice can only change by discrete amounts 0, ħ,  2ħ, … The probability f (Debye‐Waller or Lamb‐Mössbauer factor) that no lattice excitation  (zero‐phonon processes) takes place during γ‐emission or absorption. f denotes the fraction of nuclear transitions which occur without recoil. Only for this  fraction is the Mössbauer effect observable. 2 2 2mc E ER   Debye-Waller or Lamb-Mössbauer Factor 15 f increases with: * decreasing recoil energy ER * decreasing temperature T * increasing Debye temperature θD = hωD/2πkB ωD = vibrational frequency of Debye oscillator, kB = Boltzmann factor θD = a measure for the strength of the bonds between the Mössbauer atom and  the lattice Mössbauer-Experiment 16 The resonance perturbed by electric and  magnetic hyperfine interactions between the nuclei and electric and magnetic fields set up by electrons Hyperfine interactions shift and split degenerate nuclear levels resulting in  several transition lines The Mössbauer source emits a single transition line E (assume the absorber shows also only one transition line A).  E and A have slightly modified transition energies, perturbation energies 10‐8 eV  comparable to the natural linewidth, shifts the transition lines away from each other such that the overlap decreases or disappears entirely Overlap restored by the Doppler effect, i.e. by moving the absorber and the source  relative to each other Mössbauer-Experiment 17 Source and absorber are moved relative to each other Doppler velocity 57Fe :Γ0 = 4.7∙10‐9 eV, Eγ= 14400 eV, v = 0.096 mm s‐1 c = speed of light E cv 0  Mössbauer-Experiment 18 Non‐resonant background  radiation rejected 4.2 K b.p. of liquid He 1.5 K by pumping on  the liquid He vessel Several hundred channels synchronised with the vibrator synchronises a triangular voltage waveform yielding a linear Doppler velocity scale as the channel address advances controls the electro‐ mechanical vibrator the difference between the monitored signal and the reference signal drives the vibrator at the same frequency (typically 50 Hz) Mössbauer-Experiment 19 Constant acceleration Constant velocity Mössbauer-Experiment 20 the source is moved to Doppler shift the center of the emission spectrum (brown) from smaller to larger energies, relative to the center of the absorption spectrum (red ), whose center, the quantized transition energy, is fixed. The level of the transmission spectrum (blue) at each value of velocity, v, is  determined by how much the shifted emission spectrum overlaps the absorption spectrum, such that greater overlap results in reduced transmission owing to resonant absorption. The evolution of the transmission spectrum from large negative (source moving away from absorber) to large positive  values of velocity may be followed from the top to the bottom rows. Correlation of count rate with channel number and relative velocity 21 Mössbauer spectrum of metallic Fe. The  count rate is plotted as function of the  channel number Doppler velocity as a function of the channel  number Hyperfine Interactions between Nuclei and Electrons 22 Mössbauer Parameters —Electric Monopole Interaction = Isomer Shift  —Electric Quadrupole Interaction = Quadrupole Splitting EQ —Magnetic Dipole Interaction =  Magnetic Splitting EM Hyperfine Interactions 23 Electric monopole interaction (Coulombic)  between protons of the nucleus and s‐electrons penetrating the nuclear field.  Different shifts of nuclear levels A and E. Isomer shift values give information on the oxidation state, spin state, and  bonding properties, such as covalency and electronegativity. Electric quadrupole interaction between the nuclear quadrupole moment (eQ ≠ 0, I > ½) and an inhomogeneous electric field at the nucleus (EFG ≠ 0). Nuclear states split into I + ½ substates. The quadrupole splitting gives the information on oxidation state, spin state and  site symmetry. Magnetic dipole interaction between the nuclear magnetic dipole moment (μ≠ 0, I > 0) and a magnetic field (H  ≠ 0) at the nucleus. Nuclear states split into 2I+1 substates with mI= +I, +I‐1, .…,‐I The magnetic splitting gives information on the magnetic properties of the  material under study ‐ ferromagnetism, antiferromagnetism. Isomer Shift 24 The nuclear radius in the excited state is different (in 57Fe smaller) than that in  the ground state: Re ≠ Rg.  The electronic densities of all s‐electrons (1s, 2s, 3s, etc.) are different at the nuclei of the source and the absorber: ρS ≠ ρA.  The result is that the electric monopole interactions (Coulomb interactions)  are different in the source and the absorber and therefore affect the nuclear ground and excited state levels to a different extent. This leads to the measured isomer shift δ. Isomer Shift 25 δ = EA – ES = C (ρA – ρS)(Re 2 – Rg 2)     C = (2/3)πZe2. The energy levels of the ground and  excited states of a bare nucleus are  perturbed and shifted by electric monopole interactions. The shifts in the ground and excited states differ = different nuclear radii and different Coulombic interactions.  The energy differences ES and EA in the source and absorber also differ because of the different electron densities in the source and absorber material.  The individual energy differences ES and EA cannot be measured individually, a  Mössbauer experiment measures only the difference of the transition energies δ = EA – ES, isomer shift. Isomer Shift 26 The isomer shift depends directly on the s‐electron densities and are influenced indirectly via shielding by p‐, d‐, and f‐electrons which are not capable (relativistic effects) of penetrating the nuclear field. Influence on |Ψ(0)|2:  Direct = change of electron population in s‐orbitals (mainly valence s‐orbitals)  changes directly |Ψ(0)|2 Indirect = shielding by p‐, d‐, f‐electrons, increase of electron density in p‐, d‐, f‐ orbitals increases shielding effect for s‐electrons from the nuclear charge → s‐electron cloud expands, |Ψ(0)|2 at nucleus decreases. 119Sn Mössbauer Spectra 27 (Re 2 – Rg 2)  0 Compound / mm s‐1 Compound / mm s‐1 SnF4 0.36 Sn (metal) 2.50 SnO2 0.0 Sn (gray) 2.02 SnCl4 0.85 SnO 2.71 SnBr4 1.15 SnSO4 3.90 SnI4 1.55 SnF2 3.2 SnPh4 1.22 SnCl2 4.07 SnH4 1.27 SnBr2 3.93 Sn(IV)   2.00 mm s‐1 electron config. 5s0 – lower el. density at nucleus than neutral atom = negative shift Sn(II)   2.50 mm s‐1 electron config. 5s2 – higher el. density at nucleus than neutral atom as no 5p shielding = positive shift 119Sn Mössbauer Spectra 28 (Re 2 – Rg 2)  0 119Sn Mössbauer Spectra 29 (Re 2 – Rg 2)  0 PcSnX2 complexes 119Sn Mössbauer Spectra 30 10.1016/j.matchemphys.2016.07.061 JACS 1970,92,1501 IC1971,10,1553 Electron Densities at the Nucleus (r = 0) 31 The partial electron densities refer to one electron each in 1s, 2s, 3s, 4s‐orbitals.  The total s‐electron density at r = 0 is twice the sum of the partial one‐electron  contributions (all s‐orbitals are doubly occupied). Isomer Shifts of Iron Compounds 32 The most positive isomer shift occurs with iron(I)  with spin S = 3/2. The seven d‐electrons exert a very strong shielding of the s‐electrons, this reduces the s‐ electron density ρA giving a strongly negative  quantity (ρA – ρS), as (Re 2 – Rg 2)  0  for 57Fe, the isomer shift becomes strongly positive.  Strongly negative for iron(VI) with spin S = 1.  There are only two d‐electrons, the shielding effect for s‐electrons is very weak and the s‐ electron density ρA at the nucleus becomes high. Iron(II) high spin with S=2 can be easily assigned.  In other cases with overlapping δ values ambiguous assignment. Need to consider the quadrupole splitting parameter. Effect of Ligand Electronegativity 33 The electronegativity increases from I to F  In the same ordering the 4s‐electron  population decreases and the s‐electron density at the iron nucleus decreases (Re 2 – Rg 2)  0  for 57Fe the isomer shift increases from iodide to  fluoride. Fe(II) HS,  CN = 6  Second-Order Doppler Shift 34 The isomer shifts δ, i.e. the resonance peak shifts observed in  Mössbauer spectra, are composed of two terms: δ = δC + δSOD(T)  The first term is the chemical isomer shift δC which is temperature‐ independent.  The second term is the second‐order Doppler shift δSOD.  Since δSOD is T‐dependent, the observed isomer shift δ is also T‐dependent.  The second‐order Doppler shift δSOD is related to the mean‐square velocity  v2 of lattice vibrations in the direction of the γ‐ray propagation which increases  with increasing T. Accordingly, the Mössbauer resonance moves to a more  negative velocity with increasing T :  c v SOD 2 2  Electric Quadrupole Interaction Quadrupole Splitting EQ 35 Electric Quadrupole Interaction Electric quadrupole interaction occurs if at least one of the nuclear states involved possesses a quadrupole moment eQ (for I > 1/2) and if the electric field at the nucleus is inhomogeneous.  57Fe: the first excited state (14.4 keV) has a spin I = 3/2 and therefore also an electric quadrupole moment eQ. 36 Quadrupole Splitting EQ EFG is non‐zero in non‐cubic valence  electron distribution and/or in non‐cubic lattice site symmetry The precession of the quadrupole moment  vector about the field gradient axis  Splits the degenerate I = 3/2 level into two substates with magnetic spin quantum numbers mI = ±3/2 and ±1/2.  Selection rule: mI = 0, 1 The ground state I = ½  ‐no quadrupole moment  ‐unsplit ‐twofold degenerate Electric Quadrupole Interaction Quadrupole Splitting EQ 37 The energy difference between the two substates EQ is observed in the spectrum as the separation between the two resonance lines Electric Field Gradient (EFG) 38 Vij = (∂2V/∂i∂j) (i,j,k = x,y,z)  3x3 second rank EFG tensor A point charge q at a distance r = (x2+ y2+ z2)½ from the nucleus causes a  potencial V(r) = q/r at the nucleus. The electric field E at the nucleus is the negative gradient of the potential, –V, and the electric field gradient  EFG is given by: Only five Vij components are independent, because: •symmetric form of the tensor, i.e. Vij = Vji,  •Laplace: traceless tensor  Principal axes system:  |Vzz| ≥ |Vxx| ≥ |Vyy| Axial symmetry (tetragonal, trigonal) ‐ the EFG is given only by the tensor  component Vzz ,  Vxx = Vyy →  = 0 Non‐axial symmetry – the asymmetry parameter η: 0≤ η ≤ 1  = (Vxx–Vyy)/Vzz Electric Field Gradient (EFG) 39 Energy levels Two kinds of contributions to the EFG: (EFG)total = (EFG)val + (EFG)lat or in the principal axes system and η = 0: (Vzz)total = (Vzz)val + (Vzz)lat The lattice contribution(Vzz)lat = non‐cubic arrangement of the next nearest  neighbours The valence contribution (Vzz)val = anisotropic (noncubic) electron population in the orbitals Electric Field Gradient (EFG) 40 For 57Fe with Ie = 3/2, Ig = 1/2: EQ (3/2, ±3/2) = 3eQVzz/12  for I = 3/2, mI= ±3/2 EQ (3/2, ±1/2) = ‐3eQVzz/12  for I = 3/2, mI= ±1/2 the quadrupole splitting energy ΔEQ= EQ(3/2, ±3/2) ‐ EQ (3/2, ±1/2) =  eQVzz/2  (in axially symmetric systems, η= 0) [Fe(H2O)6]3+ 6A1 41 EFGlat = 0 EFGval= 0 [ML6] (Oh) z2 x2–y2 xz yz xy FeSO4•7H2O 42 [Fe(H2O)6]2+  Fe(II)‐HS, S = 2 EFGlat = 0  0 EFGval  0  0 Jahn‐Teller‐Distortion compressed elongated FeSO4•7H2O 43 K4[Fe(CN)6] 44 Fe(II)‐LS, S = 0 cubic EFGlat = 0 EFGval = 0 K4[Fe(CN)6] 45 Fe(II)‐LS, S = 0 cubic Na2[Fe(CN)5(NO)] 46 Fe(II)‐LS, S = 0  tetragonal EFGlat  0 EFGval  0 Fe(II) + NO+ or Fe(III) + NO  Na2[Fe(CN)5(NO)] 47 Fe(II)‐LS, S = 0  tetragonal Unusually large ΔEQ and too negative  for Fe(II) LS (‐0.165 mm/s)  NO withdraws electron density from dxz a dyz (‐bonding dp) Isomer Shifts and Quadrupole Splitting 48 1.0 1.5 1 -0.5 0.5 Isomer shift (mm/s) 0 2 3 4 0.0 [6] Fe(II) [6] Fe(III) [6] Fe3+ [4] Fe3+ [6] Fe2+ [4] Fe2+ [sq] Fe2+ [8] Fe2+ [5] Fe3+ [5] Fe2+ 119Sn Mössbauer Spectra and Structures of the Tin Halides 49 Magnetic Dipole Interaction 50 Magnetic Splitting EM Magnetic dipole interaction = the precession of the magnetic dipole moment vector about the axis of the magnetic field Splitting of the states I,mI into 2I + 1 substates Magnetic Dipole Interaction 51 Magnetic Splitting EM The requirements for magnetic dipole interaction ‐ the nuclear states involved possess a magnetic dipole moment  (I > ½) ‐ a magnetic field is present at the nucleus 57Fe: the ground state with I = ½ and the first excited state with I = 3/2  Selection rules: ΔI = ±1, ΔmI = 0, ±1. Magnetic Dipole Interaction 52 The energies of the sublevels : EM(mI) = – H mI / I = –gN N H mI gN = the nulcear Landé factor, N = the nuclear Bohr magneton. The separation between the lines 2 and 4 (also between 3 and 5) refers to the magnetic dipole splitting of the ground state. The separation between lines 5 and 6 (also between 1 and 2, 2 and 3, 4 and 5) refers to the magnetic dipole splitting of the excited I = 3/2 state. Internal Magnetic Field 53 A magnetic field Hint (r = 0) at the nucleus can originate in various ways: Hint = HC + HD + HL + Hext = total internal magnetic field Fermi Contact Interaction HC: Electron spin S of valence shell (e.g. S = 5/2 of Fe3+) polarizes the s‐electron density at  the nucleus: core polarisation Spin density |Ψ(0)↓|2 > |Ψ(0)↑|2 magnetic field HC ≠ 0 Spin dipolar interaction HD: The magnetic moment of the electron spin gives rise to dipolar interaction  with the nucleus and causes a field at r = 0. Internal Magnetic Field 54 Orbital dipolar interaction HL: Electrons with orbital moment L ≠ 0 give rise to an orbital magne c moment  accompanied by a magnetic field HL= ‐2 μB : expectation value of orbital angular momentum. Externally applied field Hext: By applying an external magnetic field of known size and direction one can determine the size and the direction of the intrinsic magnetic field of the material under investigation. Combined Magnetic Dipole and Electric Quadrupole Interactions 55 Combined Magnetic Dipole and Electric Quadrupole Interactions 56 Magnetic dipole interaction and electric quadrupole interaction may be present in  a material simultaneously (together with the electric monopole interaction which is always present).  Relatively weak quadrupole interaction The nuclear sublevels I,mI arising from magnetic dipole splitting are additionally shifted by the quadrupole interaction energies EQ(I,mI) The sublevels of the excited I = 3/2 state are no longer equally spaced. The shifts by EQ are upwards or downwards depending on the direction of the EFG.  This enables one to determine the sign of the quadrupole splitting parameter ΔEQ. The quadrupole shift parameter  depends on the canting angle  of the spins with respect to the electric field gradient (EQ) axis [111]  = EQ(3 cos2  ‐ 1)/2 and thus yields values with opposite sign for AF ( = 0°)  and WF ( = 90°) states. Magnetic Dipole Interaction 57 Fe3(CO)12 58 [Fe3(3-O)(OAc)6(H2O)3] 59 Variable‐temperature 57Fe Mössbauer spectra [Fe3(3-O)(OAc)6(3-Et-py)3]S 60 Variable‐temperature 57Fe MÖssbauer spectra S = CH3CCl3 CH3CCl3 is valence‐trapped at low temperature, increasing temperature leads to valence  detrapping near room temperature. The ratio of the area fractions of Fe(III) to  Fe(II) is close to 2 at low temperatures. [Fe3(3-O)(OAc)6(3-Et-py)3]S 61 Variable‐temperature 57Fe MÖssbauer spectra S = C6H6 CH3CN increasing temperature leads to valence detrapping near room temperature.  The lattice packing controls valence de/trapping.  C6H6  possesses a stack type  structure with strong intermolecular interactions due to overlapping pyridine ligands.  CH3CN and CH3CCl3 arelayered. CH3CN C6H6 is valence‐ trapped from 120  to 298 K on the Mössbauer time scale (given by the lifetime of the nuclear excited state).  γ-Radiolysis of FeSO4· 7H2O (300 K) 62 Corrosion Products 63 Corrosion of α‐Iron in H2O/SO2 atmosphere at 300 K Corrosion product is β‐FeOOH Iron(III) Oxides 64 Iron(III) Oxides 65 alpha‐Fe2O3 measured at 260 K  near Morin transition temperature Hematite At < 260 K is antiferromagnetic (AF) with the  spins oriented along the electric field gradient  axis.  At Morin temperature (TM), around 260 K, a  reorientation of spins by about 90°, the spins become slightly canted to each other (by 5°),  causing the destabilization of their perfect  antiparallel arrangement = weak (parasitic)  ferromagnetism between Morin and Neel  temperature (TN). Above the Neel temperature of 950 K, hematite  loses its magnetic ordering and is paramagnetic. Iron(III) Oxides 66 Maghemite Mössbauer spectrum of a well‐ crystallized gamma‐Fe2O3 at 4 K in  an external field of 6 T. Iron in French Red Wine 67