A complete solution to the Mössbauer problem, all in one place

Similar documents
3. Perturbed Angular Correlation Spectroscopy

Hyperfine interaction

CHAPTER II. The Angular and Polarization Distributions of the Mossbauer Radiation 3/2 + 1/2 + in Single Crystals *

Electric and magnetic multipoles

Physics 221A Fall 2005 Homework 11 Due Thursday, November 17, 2005

Physics 221A Fall 1996 Notes 21 Hyperfine Structure in Hydrogen and Alkali Atoms

PH 451/551 Quantum Mechanics Capstone Winter 201x

Tutorial Session - Exercises. Problem 1: Dipolar Interaction in Water Moleclues

Physics 115C Homework 2

Total Angular Momentum for Hydrogen

Time Independent Perturbation Theory Contd.

Lecture 4 Quantum mechanics in more than one-dimension

Conclusion. 109m Ag isomer showed that there is no such broadening. Because one can hardly

NMR: Formalism & Techniques

Doping-induced valence change in Yb 5 Ge 4 x (Sb, Ga) x : (x 1)

Fun With Carbon Monoxide. p. 1/2

Lecture 4 Quantum mechanics in more than one-dimension

Addition of Angular Momenta

Final Exam Tuesday, May 8, 2012 Starting at 8:30 a.m., Hoyt Hall Duration: 2h 30m

Phys 622 Problems Chapter 5

The Hydrogen Atom. Dr. Sabry El-Taher 1. e 4. U U r

An Introduction to Hyperfine Structure and Its G-factor

3 Angular Momentum and Spin

The tensor spherical harmonics

1 Supplementary Figure

Quantum Mechanics: The Hydrogen Atom

Partial Dynamical Symmetry in Deformed Nuclei. Abstract

Gamma-ray decay. Introduction to Nuclear Science. Simon Fraser University Spring NUCS 342 March 7, 2011

6.1 Nondegenerate Perturbation Theory

4/21/2010. Schrödinger Equation For Hydrogen Atom. Spherical Coordinates CHAPTER 8

Spin Dynamics Basics of Nuclear Magnetic Resonance. Malcolm H. Levitt

Alkali metals show splitting of spectral lines in absence of magnetic field. s lines not split p, d lines split

Angular momentum and spin

MASSACHUSETTS INSTITUTE OF TECHNOLOGY 5.76 Modern Topics in Physical Chemistry Spring, Problem Set #2

Physics 218 Polarization sum for massless spin-one particles Winter 2016

Relativistic corrections of energy terms

Physics 228 Today: Ch 41: 1-3: 3D quantum mechanics, hydrogen atom

Nuclear models: Collective Nuclear Models (part 2)

Optical pumping and the Zeeman Effect

Fine Structure of the metastable a 3 Σ u + state of the helium molecule

LECTURES ON QUANTUM MECHANICS

A.1 Alkaline atoms in magnetic fields

14. Structure of Nuclei

(relativistic effects kinetic energy & spin-orbit coupling) 3. Hyperfine structure: ) (spin-spin coupling of e & p + magnetic moments) 4.

Prob (solution by Michael Fisher) 1

In this lecture, we will go through the hyperfine structure of atoms. The coupling of nuclear and electronic total angular momentum is explained.

Forbidden Electric Dipole Transitions in the Hydrogen Molecular Ion First Estimates

8 NMR Interactions: Dipolar Coupling

Hyperfine interactions Mössbauer, PAC and NMR Spectroscopy: Quadrupole splittings, Isomer shifts, Hyperfine fields (NMR shifts)

Supplementary Information: Electrically Driven Single Electron Spin Resonance in a Slanting Zeeman Field

Chem 442 Review for Exam 2. Exact separation of the Hamiltonian of a hydrogenic atom into center-of-mass (3D) and relative (3D) components.

Supplementary Information

Mossbauer Effect. Ahmad Ali Ohyda. 1 Ahmad Faraj Abuaisha. 2 Abu-Bakr Mohammad Alrotob. 3 Basher M. Ismail. 4. Abstract

Quantum Theory of Many-Particle Systems, Phys. 540

Rotations in Quantum Mechanics

Problem 1: Spin 1 2. particles (10 points)

Physics 221A Fall 2017 Notes 25 The Zeeman Effect in Hydrogen and Alkali Atoms

One-electron Atom. (in spherical coordinates), where Y lm. are spherical harmonics, we arrive at the following Schrödinger equation:

Lecture 4: Nuclear Energy Generation

Rotations and vibrations of polyatomic molecules

and for absorption: 2π c 3 m 2 ˆɛˆk,j a pe i k r b 2( nˆk,j +1 February 1, 2000

Chapter 8 Magnetic Resonance

H atom solution. 1 Introduction 2. 2 Coordinate system 2. 3 Variable separation 4

20 The Hydrogen Atom. Ze2 r R (20.1) H( r, R) = h2 2m 2 r h2 2M 2 R

CHAPTER 6: AN APPLICATION OF PERTURBATION THEORY THE FINE AND HYPERFINE STRUCTURE OF THE HYDROGEN ATOM. (From Cohen-Tannoudji, Chapter XII)

1.6. Quantum mechanical description of the hydrogen atom

Magentic Energy Diagram for A Single Electron Spin and Two Coupled Electron Spins. Zero Field.

(n, l, m l ) 3/2/2016. Quantum Numbers (QN) Plots of Energy Level. Roadmap for Exploring Hydrogen Atom

Quantum Physics II (8.05) Fall 2002 Assignment 11

The concept of free electromagnetic field in quantum domain

Quantum Mechanics FKA081/FIM400 Final Exam 28 October 2015

Problem Set #4: 4.1,4.7,4.9 (Due Monday, March 25th)

Physics 4022 Notes on Density Matrices

Quantization of the E-M field

Atomic and molecular physics Revision lecture

16.1. PROBLEM SET I 197

Magnetic Resonance Spectroscopy ( )

Multi-Electron Atoms II

ECEN 5005 Crystals, Nanocrystals and Device Applications Class 20 Group Theory For Crystals

P3317 HW from Lecture 7+8 and Recitation 4

Quantum Physics III (8.06) Spring 2007 FINAL EXAMINATION Monday May 21, 9:00 am You have 3 hours.

POEM: Physics of Emergent Materials

G : Quantum Mechanics II

Scattering cross-section (µm 2 )

IV. Electronic Spectroscopy, Angular Momentum, and Magnetic Resonance

Massachusetts Institute of Technology Physics Department

Spin Interactions. Giuseppe Pileio 24/10/2006

Nuclear Shell Model. Experimental evidences for the existence of magic numbers;

Fundamentals of Spectroscopy for Optical Remote Sensing. Course Outline 2009

Notes on Quantum Mechanics

Appendix: SU(2) spin angular momentum and single spin dynamics

Part II Particle and Nuclear Physics Examples Sheet 4

Optical Pumping in 85 Rb and 87 Rb

Diatomics in Molecules - DIM

Atomic Physics 3 rd year B1

3. Quantum Mechanics in 3D

6 NMR Interactions: Zeeman and CSA

Quantum Physics II (8.05) Fall 2002 Assignment 12 and Study Aid

5.80 Small-Molecule Spectroscopy and Dynamics

Printed from file Manuscripts/Stark/stark.tex on February 24, 2014 Stark Effect

Transcription:

Hyperfine Interact 006) 70:9 04 DOI 0.007/s07-006-9467- A complete solution to the Mössbauer problem, all in one place C. J. Voyer D. H. Ryan Published online: December 006 Springer Science + Business Media B.V. 006 Abstract We present a full solution to the general combined interactions static Mössbauer problem that is easily generalized to any Mössbauer isotope, and applies for M, E, and E transitions as well as combined M E transitions. Explicit expressions are given for both powder and single crystal samples. Key words Mössbauer spectrum calculation full Hamiltonian diagonalization combined hyperfine interactions Introduction While first-order perturbation approaches or interpolation schemes are adequate for fitting the majority of simple Mössbauer spectra, there are numerous cases for which these approaches are invalid. If contributions from magnetic and electrostatic interactions are of comparable magnitude or do not share a common quantization axis, a full solution to the nuclear Hamiltonian must be used to fit the spectrum. Although functional codes have been published in the past [ ], in the current computational climate, where not only programming languages compete, but where computional packages that include programmable options Matlab, Mathematica...) are also used, an explicit and adaptable mathematical solution is more useful. Most calculations given are either not worked out completely [, ], or are far too involved mathematically to be easily generalized and implemented into simple programs [4]. In order to promote the more widespread implementation of full solution Mössbauer fitting codes, we present here a complete solution that we have developed. This solution can be generalized to cover any accessible Mössbauer isotope and is cast in a form that can be readily coded into a fitting routine. C. J. Voyer D. H. Ryan B) Centre for the Physics of Materials and Physics Department, McGill University, 600 University Street, Montréal, QC HA T8, Canada e-mail: dominic@physics.mcgill.ca

9 C.J. Voyer, D.H. Ryan The Mössbauer problem The Mössbauer problem consists of computing a Mössbauer spectrum given a set of hyperfine parameters. The solution can then be used in a least squares refinement of an experimentally obtained Mössbauer spectrum. The line positions depend on the splitting of the ground and excited states and their relative energy difference, and can be derived from the eigenvalues of the system Hamiltonian. Line intensities depend on the coupling of the two angular momentum states involved in the transition, and can be obtained from the eigenvectors of the Hamiltonian combined with appropriate vector coupling coefficients []. In the absence of hyperfine fields, the ground and excited states are degenerate, and a single line spectrum is observed. When an electric field gradient, or a magnetic field, or both are present, the degeneracy is lifted and multi-line spectra are obtained as shown in Figure. In the static model, the interaction Hamiltonian H is given by [6]: H = H + H C + H M + H Q +... ) H represents all terms for the atom not including hyperfine interactions and does not influence the energy difference between the ground and excited states of the nucleus. Having no impact on the Mössbauer spectrum, it is ignored. H C represents the Coulombic interactions between the nucleus and the electrons, i.e. the electric monopole term, and has the effect of shifting all excited state energy levels with respect to all ground state energy levels, and is called the isomer shift IS). The IS will have the effect of shifting the entire Mössbauer spectrum either up or down in velocity. Other effects can also shift the entire spectrum by some constant velocity value, and so for fitting purposes we consider the more general centre shift CS), which can be added to any solution once the more complex calculations are completed. Finally, H M and H Q, the magnetic dipole and electric quadrupole contributions, are the terms that lift the energy level degeneracy, and are the nontrivial part of the computation. All higher order terms are ignored [6].. Magnetic dipole interactions For a nucleus in a state with a gyromagnetic ratio g, and subject to a hyperfine magnetic field B, H M can written as [7]: H M = gμ N B Î ) where Î is the total spin operator and μ N is the nuclear magneton. For simplicity, we may choose the z-axis to be the direction of the field and get: H M = gμ N BÎ z ) where B is the magnitude of the field and Î z is the z component of the angular momentum operator. For both the excited and ground states, the m states are eigenstates of this Hamiltonian and thus the eigenvalues are: and the eigenvector matrix is simply identity. E M = gμ N Bm 4)

A complete solution to the Mössbauer problem, all in one place 9 ±/ ±/ I=/ +/ +/ / / ±/ I=/ / +/ Figure Sketch of energy level degeneracy lifting in the 7 Fecase.Inthecentre we see the degenerate excited and ground states, on the left the excited state has the degeneracy partially lifted when an electric field gradient is present, and on the right when a magnetic field is present, we see the degeneracy completely lifted. The arrows drawn are the allowed transitions given the M multipolarity.. Electric quadrupole interactions The nuclear energy level degeneracy is also lifted by the quadrupole moment of the probe nucleus interacting with an electric field gradient. The electric field gradient is a traceless tensor: V ij = V/ x i x j ) x i, x j = x, y, z). ) The coordinate system can always be chosen such that the tensor is diagonal and traditionally the z-axis is chosen so that V zz is its largest component. Since the tensor is traceless, V xx + V yy + V zz = 0, only two parameters are required to specify the tensor completely. We define η = Vxx Vyy V zz, using the convention V zz V yy V xx to ensure that 0 η, and the tensor can be specified by V zz and η. For a nucleus with an electric quadrupole moment Q, spini, subject to an electric field gradient whose principal axis value is V zz,withanasymmetryη, the Hamiltonian is [8]: H Q = eqv zz [Îz 4II ) II + ) + η ] Î + Î ) 6) where e is the charge of the proton and Î + and Î are the ladder operators as defined by Sakurai [9]. If η = 0 the m states are energy eigenstates and the energies are E m = eqv zz [ 4II ) m II + ) ], which in turn yields a degeneracy in the energy levels as only the magnitude of m contributes to the energy. Complications arise when η = 0 since in this case the m states are no longer energy eigenstates. The Hamiltonian is then non-diagonal and the eigenvectors need to be computed explicitly and a full diagonalization is necessary. A fitting routine needs to compute a full spectrum many times in order to complete a pattern refinement. Available computing power used to limit solution possibilities, and strategies were devised to avoid repeated calculation of the full solution. Exact analytic solutions for the energy levels exist for the I = case [0], and the I = case

94 C.J. Voyer, D.H. Ryan [], however, when the dimensions of the Hamiltonian get larger the mathematics of an analytic solution becomes prohibitively difficult. Shenoy et al. [] generated a polynomial interpolation scheme, where the line positions and intensities have been computed exactly for incremental values of η and then fit with a polynomial whose coefficients were made available. Finally the perturbative approach is useful if one of the interactions is small compared to the other. Today, however, computing power limitations for this type of calculation no longer exist, and the only remaining barrier to a full solution is the mathematics involved. The ideal situation is to have a solution to the Mössbauer problem that can be implemented easily as a numerical routine that yields exact line positions and intensities given a set of hyperfine parameters, with no limitations on the values of these parameters. This solution could then be incorporated into a least squares fitting routine and refine experimentally obtained data. Numerical diagonalization is longer a limitation. We can solve the general static Mössbauer problem exactly, and have a solution that can cover the most complex cases, as well as covering simple cases, that can be easily implemented, operate efficiently and satisfy all fitting needs. Combined interactions The most general physical situation that can be considered is the case of combined magnetic dipole and electric quadrupole interactions. In the previous cases considered, the optimal choice of quantization axis was evident. Here we have two logical choices: the magnetic field direction, or the principal axis of the electric field gradient, V zz. Typically, V zz is chosen as the quantization axis and the schematic setup is shown in Figure. Note that this choice of axes is valid even if V zz = 0, and none of the computations change. From ), the Hamiltonian operator for this system becomes [8]: H = H Q + H M 7) where H Q does not change from 6), but H M needs to be rotated away from the z-axis and can be written following the convention of Figure as: H M = gμ N B hf [Î z cos θ + ] Î + e iφ + Î e iφ ) sin θ 8) This Hamiltonian will not, in general, be diagonal in the m basis. To compute the energy eigenvalues and eigenvectors we need to write the Hamiltonian in matrix form. The expression for the matrix depends on the total spins I g and I e of the ground and excited states. Once the Mössbauer transition is chosen, and I g and I e are thereby fixed, we can write the Hamiltonian for the ground and excited states, H g and H e. For example, if we are interested in a transition as in 7 Fe, 9 Sn...) or a as in 89 Os, 97 Au...) transition, if we set A = eqvzz where Q is the 4II ) appropriate value for the quadrupole moment of the nucleus in the I = state, and

A complete solution to the Mössbauer problem, all in one place 9 Figure Schematic of the φ and θ angles, V zz and B hf as they appear in the Hamiltonian of 7). V zz θ B hf φ V yy V xx α = g μ N B hf,whereg is the appropriate value for the gyromagnetic factor of the nucleus in the I = state, we may write matrices constructed as described in []): H / = and setting β = g / μ N B hf : A α cos θ α sin θe iφ Aη 0 α sin θeiφ A α cos θ α sin θe iφ Aη Aη α sin θe iφ A + α cos θ 0 H / = α sin θe iφ Aη α sin θeiφ A + α cos θ β cos θ β sin θe+iφ + β cos θ β sin θe iφ ) 9) 0) The I = 0 case is of particular mathematical interest because the Hamiltonian in 7) is null. This case is treated by considering the state to have a unidimensional eigenvector in the m basis, and an energy eigenvalue of 0. So if we are interested in the 0 transition 70 Yb, 66 Er...), the only Hamiltonian that needs to be explicitly computed is the I = excited state Hamiltonian here β = g μ N B hf ): 6A β cos θ βe iφ sin θ A 6η 0 0 βe iφ sin θ A β cos θ 6 β sin θe iφ Aη 0 H = A 6η β 6 eiφ sin θ 6A 6 β sin θe iφ B 6η 0 Aη 6 β sin θeiφ A + β cos θ β sin θe iφ 0 0 A 6η β sin θe iφ 6A + β cos θ With the excited and ground state Hamiltonians now analytically expressed, a routine may compute the components, given a set of hyperfine parameters, then proceed to the numerical diagonalization, and finally compute the line positions and intensities. )

96 C.J. Voyer, D.H. Ryan Many options are available for the numerical diagonalization. We have used the tred and tqli routines outlined in Numerical Recipes [4, ]. The Matlab and Mathematica software packages also offer diagonalization functions. To avoid working with complex numbers Numerical Recipes suggests a scheme by which the matrix can be kept real, however it requires doubling the dimensionality of the problem. After diagonalization is achieved, the problem is then reduced back to its original dimension by selecting half of the eigenvalues and eigenvectors. The selection must be done with care since the ordering of the diagonalization output depends on the set of non-zero hyperfine parameters. The selection must ensure a set of mutually orthogonal eigenvectors is obtained. From these routines, we obtain numerical values for our eigenvectors e n as well as the amplitudes n m and the eigenvalues of the system. For a state with total spin I and corresponding eigenvalues e n, for example: m = I n en m = I n en e n = m = I n en. ) At this stage we have the eigenvalues which will give us the line positions, and the m n factors which will give us the intensity of each line. We have thus obtained all the necessary information from the Hamiltonians of the excited and ground states, and may now proceed to calculate the exact line positions and intensities.. Line position calculation The energy shift detected by Mössbauer spectroscopy is the difference between the energies of the excited state Hamiltonian E ne and the ground state Hamiltonian E ng. For a given transition between the levels E ng of the ground state and E ne of the excited state, we obtain a line position E n e g): E n e g) = E ne E ng ) The constant offset CS mentioned in Section can then be added, and the final result is completely general for all values of the hyperfine parameters. We note that for the specific case of the 0 transition 70 Yb), this reduces to: E n 0) = E n + CS 4) The number of lines observed will depend on the dimensionality of both H e and H g, and also on the multipolarity of the transition i.e. the quantum mechanical selection rules that apply in each situation.. Line intensity calculation The relative probability of each transition s occurrence will depend on the coupling of the two angular momentum states involved in the transition []. Several methods exist for calculating the relative intensity of a line [,, 8, 6]. These treatments fall short however, in that they either treat a single case [8], do not treat the E case [6],

A complete solution to the Mössbauer problem, all in one place 97 do not give an explicit and easily generalizable solution [,, 8, 6], or do not cover the single crystal case [,, 8]. We chose to treat the relative intensity of a line in a spectrum as the normalized power absorbed by the corresponding transition. We write the multipole electric field vector E as [7]: Eϕ, ϑ) = l,m i) l+ [ a E l, m)x m l r + a M l, m)xl m ] ) where ϑ and ϕ are the polar angles describing the direction of emission of the γ, along r, with respect to our chosen axis of quantization not to be confused with φ and θ defined in Section ); Xl m are the vector spherical harmonics, and will be derived later Section.); and the a E l, m) and a M l, m) specify the amounts of electric multipole and magnetic multipole fields, and are determined by the source and by boundary conditions derived in Section.4)[8]. Since we are only concerned with relative intensities, factors common to all transitions can be normalized out and so are omitted from ). The power absorbed i.e. the intensity of a line) will be E.E. This method simply and naturally introduces the vector spherical harmonics needed, and leads to a straightforward single crystal intensity as well as a powder average intensity. This method is preferred because of its simple computations and adaptability to different transitions the experimentalist may face. The three transition multipolarities encountered in Mössbauer spectroscopy are the M, E and E cases. Only the M and E are ever combined, and the E case is completely analogous to the M case and requires no new information, thus for the purposes of this computation, we will not address the E case. From ) we get for the M case, and for the E case we get E M ϕ, ϑ) = E E ϕ, ϑ) = + m= + m= a M, m).x m l 6) i. [a E, m)].x m l r) 7) To proceed further, we need to express E more explicitly, thus calculating the vector spherical harmonics X m l, and the amounts of electric and magnetic multipole fields a E l, m) and a M l, m).. The X m l The vector spherical harmonics, X m l, can be generated from [8] X m l = ll + ) ˆLY m l ϑ, ϕ) 8)

98 C.J. Voyer, D.H. Ryan where, ˆL = i ˆϑ sin ϑ ϕ ˆϕ ) ϑ 9) We compute the vector spherical harmonics needed, X = ) e iϕ ˆϑ i cos ϑe iϕ ˆϕ 0) X 0 = i sinϑ ˆϕ ) X = ) e iϕ ˆϑ + i cos ϑe iϕ ˆϕ ) X = ) sin ϑe iϕ ˆϑ i sin ϑ cos ϑe iϕ ˆϕ ) X = cos ϑe iϕ ˆϑ i [ cos ϑ sin ϑ ] ) e iϕ ˆϕ 4) X 0 = ) 6i cos ϑ sin ϑ ˆϕ ) X = cos ϑe iϕ ˆϑ + i [ cos ϑ sin ϑ ] ) e iϕ ˆϕ 6) X = ) sin ϑe iϕ ˆϑ i sin ϑ cos ϑe iϕ ˆϕ 7) We note that in spherical coordinates ˆϑ ˆr =ˆϕ and that ˆϕ ˆr = ˆϑ, such that for the X m of the E part we get X ˆr = ) sin ϑe iϕ ˆϕ + i sin ϑ cos ϑe iϕ ˆϑ X ˆr = X 0 ˆr = X ˆr = X ˆr = 4 cos ϑe iϕ ˆϕ + i [ cos ϑ sin ϑ ] ) e iϕ ˆϑ 8) 9) 6i cos ϑ sin ϑ) ˆϑ 0) cos ϑe iϕ ˆϕ i [ cos ϑ sin ϑ ] ) e iϕ ˆϑ ) sin ϑe iϕ ˆϕ + i sin ϑ cos ϑe iϕ ˆϑ π ) )

A complete solution to the Mössbauer problem, all in one place 99.4 The al, m) For this particular problem, the a E L, m) and a M L, m) coefficients have the same analytic expression [6], al, m) = m e n e m g n g I g Lm g m I e m e ) m e m g=m where n g and n e are the energy eigenstates of the ground and excited states respectively, and m e and m g are the Î z eigenstates for the excited and ground states respectively, note that m e n e and m g n g are simply the components of the eigenvectors of H e and H g, respectively), and I g Lm g m I e m e are Clebsch-Gordan coefficients where L is the multipolarity of the transition for M, L = and for E, L = ), m = m e m g, I e is the total spin of the excited state and I g is the total spin of the ground state. For the 0 transition, all Clebsch-Gordon coefficients are exactly since I g =0. But if the spins for both the excited and ground states are non-zero, the Clebsch- Gordon coefficients are non-trivial see Table I for an example). Note that the I g Lm g m I e m e coefficients are first converted using the relation [9]: I I g Lm g m I e m e = ) Ig mg e + I e I g ) + I ei g m e m g Lm 4) Once an expression for the Clebsch-Gordon coefficients is obtained we can move on to calculating the al, m) explicitly. Again, for the case of a / / transition we get, 4 a, ) = / n e / n g a, ) = / n e / n g / n e / n g a, 0) = / n e / n g / n e / n g a, ) = / n e / n g / n e / n g 4 a, ) = / n e / n g a, ) = / n e / n g + / n e / n g a, 0) = / n e / n g + / n e / n g a, ) = / n e / n g + / n e / n g

00 C.J. Voyer, D.H. Ryan Table I Clebsch-Gordon coefficients pertinent to the / / M and/or E solution E M m g m e m m gm m e m gm m e + + + + + 0 + + 4 0 + 4 + 0 + + 0 where the a, m) are for the E multipolarity case, and the a, m) are for the M multipolarity case. For the 0 case with E multipolarity however, the expressions become much simpler, a, ) = n e a, ) = n e a, 0) = 0 n e a, ) = n e a, ) = n e The vectors in 6)and7) now have numerical values. We can now move on to calculating the intensity of each line.. M transitions We first consider the case of an M transition 7 Fe, 9 Sn...).Theelectricfieldvector can be written as: E M ϕ, ϑ) = 4 π [ a, )e iϕ + a, )e iϕ] ˆϑ [ + i a, )e iϕ cos ϑ + a, 0) ] ) sinϑ a, )e iϕ cos ϑ ˆϕ )

A complete solution to the Mössbauer problem, all in one place 0 Noting that ˆϕ.ˆϑ = 0, the intensity of a line can be written as: E M.E M = a, ) + cos ϑ) + a, 0) sin ϑ 6π + a, ) + cos ϑ) + R [ a, )a, )e iϕ] + R [a, )a, 0) sinϑ cos ϑe iϕ] R [ a, )a, ) cos ϑe iϕ] [ R a, 0)a, ) sinϑ cos ϑe iϕ] ) 6) Equation 6 is the intensity for a line in a single crystal Mössbauer absorption with a γ absorption that makes angles ϕ and ϑ with respect to the principal axis of the electric field gradient V zz.the factor can be ignored as it is the same for all 6π transitions. Most samples used are polycrystalline, in which the intensity needs to be integrated over all directions. The intensity for a powder sample becomes: π π ϑ=0 ϕ=0 E M.E M sin ϑdϕdϑ = a, ) + a, 0) + a, ) 7) We note that although one of the most cited solutions of the Mössbauer problem Kündig s solution of the 7 Fe problem [8]) yields a slightly different expression, his expression can be simplified to obtain 7), providing an independent verification that our process is correct..6 E transitions For an E transition the selection rules are less restrictive and far more transitions are allowed. The electric field vector can be written as: E E = 4 [ a, ) sin ϑ cos ϑe iϕ a, ) [ cos ϑ sin ϑ ] e iϕ π + a, 0) 6cosϑ sin ϑ + a, ) [ cos ϑ sin ϑ ] e iϕ a, ) sin ϑ cos ϑe iϕ] [ ˆϑ + i a, ) sin ϑe iϕ + a, ) cos ϑe iϕ + a, ) cos ϑe iϕ a, ) sin ϑe iϕ] ) ˆϕ 8)

0 C.J. Voyer, D.H. Ryan The intensity is then: E E.E E = a, ) sin ϑ+ cos ϑ)+ a, ) [ cos ϑ sin ϑ ] ) +cos ϑ 6π + 6 a, 0) cos ϑ sin ϑ + a, ) [ cos ϑ sin ϑ ] ) + cos ϑ + a, ) sin ϑ cos ϑ + sin ϑ ) + R { a, )a, ) sin ϑ cos ϑ [ cos ϑ sin ϑ ] e iϕ} R {a, )a, 0) 6cos ϑ sin ϑe iϕ} R { a, )a, ) sin ϑ cos ϑ [ cos ϑ sin ϑ ] e iϕ} + R { a, )a, ) cos ϑ sin ϑe 4iϕ} R {a, )a, 0) 6sinϑ cos ϑ [ cos ϑ sin ϑ ] e iϕ} R {a, )a, ) [ cos ϑ sin ϑ ] } e iϕ + R { a, )a, ) sin ϑ cos ϑ [ cos ϑ sin ϑ ] e iϕ} + R {a, 0)a, ) 6sinϑ cos ϑ [ cos ϑ sin ϑ ] e iϕ} R {a, 0)a, ) 6cos ϑ sin ϑe iϕ} R { a, )a, ) sin ϑ cos ϑ [ cos ϑ sin ϑ ] e iϕ} + R [ a, )a, ) cos ϑ sin ϑe iϕ] + R [ a, )a, ) sin ϑ cos ϑe iϕ] R [ a, )a, ) sin ϑe 4iϕ] + R [ a, )a, ) cos ϑe iϕ] R [ a, )a, ) sin ϑ cos ϑe iϕ] R [ a, )a, ) cos ϑ sin ϑe iϕ] ) 9) Again, as was done in Section., we integrate over all directions to obtain the intensities for the powder average case and obtain, π ϑ=0 π ϕ=0 E.E sin ϑdϑdϕ = a, ) + a, ) + a, 0) + a, ) + a, ) 40)

A complete solution to the Mössbauer problem, all in one place 0 a result which is analogous to the M case and completes the list of results necessary to compute the exact line intensity of any static combined interactions Mössbauer spectrum. For the 0 case, the intensity expression is greatly simplified: π ϑ=0 π ϕ=0 E.E sin ϑdϑdϕ = i + i + 0 i + i + i ) = i m m i making the intensity computation unnecessary. m = 4).7 Mixed M E transitions Certain Mössbauer isotopes have a mixture of M and E transitions in nonnegligible amounts e.g. 97 Au where E/M=0.). If we define X = intensity expression is: E E+M,the E.E = X ) E M.E M + X E E.E E 4) Thus the single crystal intensity for a combined E-M transition is the weighted sum of 6)and9), and the powder intensity is the weighted sum of 7)and40). 4 Adapting the solution to other isotopes When faced with an isotope not treated in the present text, the researcher may simply apply the same general strategy used here, and indeed the more complex expressions still hold. 6), 7), 9)and40) do not change. A new isotope brings new values for the total spin of the excited and ground states and different values for the quadrupole moments and gyromagnetic ratios for the excited and ground states. The spectra obtained from different isotopes will behave differently but the strategies outlined in this paper can be easily applied:. Set up the matrices H e and H g, as was done in 9), 0)and).. Set up the chosen diagonalization routine and store proper set of eigenvalues and eigenvectors. Calculate the sum in ) using the appropriate Clebsch-Gordon coefficients. The number of terms is set by: m e m g = m, andthe m n are the components of the eigenvectors as shown in ). 4. The intensities are computed using the relevant equation from 6),7), 9), 40), or 4), depending on the transition multipolarity M,E, or M E) and whether the experiment is on a single crystal or powder sample.. The line positions are calculated using ). This strategy is completely general for any static combined interactions case, for any Mössbauer isotope.

04 C.J. Voyer, D.H. Ryan References. Gabriel, J.R., Ruby, S.L.: Nucl. Instrum. Methods 6, 96). Kündig, W.: Nucl. Instrum. Methods 7, 6 969). Chipaux, R.: Comput. Phys. Commun. 60, 40 990) 4. Szymański, K.: J. Phys., Condens. Matter, 749 000). Gibb, T.C., Greenwood, N.N.: In: Mössbauer Spectroscopy, p.66. Chapman and Hall, London, UK 97) 6. Gibb, T.C., Greenwood, N.N.: In: Mössbauer Spectroscopy, p.46. Chapman and Hall, London, UK 97) 7. Gibb, T.C., Greenwood, N.N.: In: Mössbauer Spectroscopy, p.60. Chapman and Hall, London, UK 97) 8. Kündig, W.: Nucl. Instrum. Methods 48, 9 967) 9. Sakurai, J.J.: In: Modern Quantum Mechanics, nd edn., p.9. Addison-Wesley, New York 994) 0. Gibb, T.C., Greenwood, N.N.: In: Mössbauer Spectroscopy, p.6. Chapman and Hall, London, UK 97). Gerdau, E., Wolf, J., Winkler, H., Braunsfurth, J.: Proc. Roy. Soc. Ser. A, 97 06 969). Shenoy, G.K., Dunlap, B.D.: Nucl. Instrum. Methods 7, 8 969). Sakurai, J.J.: In: Modern Quantum Mechanics, nd edn., p.0. Addison-Wesley, New York 994) 4. Press, W.H., Tukolsky, S.A., Vetterling, W.T., Flannery, B.P.: In: Numerical Recipes in C, nd edn., p.474. Cambridge University Press, New York 999). Press, W.H., Tukolsky, S.A., Vetterling, W.T., Flannery, B.P.: In: Numerical Recipes in C, nd edn., p.480. Cambridge University Press, New York 999) 6. Housley, R.M., Grant, R.W., Gonser, U.: Phys. Rev. Lett. 78, 4 969) 7. Jackson, J.D.: In: Classical Electrodynamics, rd edn., p.47. Wiley, Hoboken, NJ 999) 8. Jackson, J.D.: In: Classical Electrodynamics, rd edn., p.4. Wiley, Hoboken, NJ 999) 9. Edmonds, A.R.: In: Angular Momentum in Quantum Mechanics, p.4. Princeton University Press, Princeton, New Jersey 97)