Effective field theory for deformed odd-mass nuclei
Abstract
We develop an effective field theory (EFT) for deformed odd-mass nuclei. These are described as an axially symmetric core to which a nucleon is coupled. In the coordinate system fixed to the core the nucleon is subject to an axially symmetric potential. Power counting is based on the separation of scales between low-lying rotations and higher-lying states of the core. In leading order, core and nucleon are coupled by universal derivative terms. These comprise a covariant derivative and gauge potentials which account for Coriolis forces and relate to Berry-phase phenomena. At leading order, the EFT combines the particle-rotor and Nilsson models. We work out the EFT up to next-to-leading order and illustrate the results in 239Pu and 187Os. At leading order, odd-mass nuclei with rotational band heads that are close in energy and differ by one unit of angular momentum are triaxially deformed. For band heads that are well separated in energy, triaxiality becomes a subleading effect. The EFT developed in this paper presents a model-independent approach to the particle-rotor system that is capable of systematic improvement.
I Introduction
In the last two decades, ideas based on effective field theory (EFT) and on the renormalization group have exerted a strong influence on nuclear-structure theory Bedaque and van Kolck (2002); Bogner et al. (2003); Epelbaum et al. (2009); Bogner et al. (2010); Grießhammer et al. (2012); Hammer et al. (2017). These ideas have led to model-independent approaches to nuclear interactions, currents, and nuclear spectra, to a new understanding of resolution-scale and scheme dependences in theoretical calculations Bedaque et al. (1999); Furnstahl and Hammer (2002); Anderson et al. (2010), and to quantitative estimates of theoretical uncertainties Schindler and Phillips (2009); Furnstahl et al. (2015). EFT exploits a separation of scale between the low-energy phenomena of interest and the excluded high-energy aspects. Thus, EFT can also be used to describe low-lying collective nuclear excitations such as rotations Papenbrock (2011); Papenbrock and Weidenmüller (2014); Papenbrock and WeidenmÃŒller (2015); Papenbrock and Weidenmüller (2016); Chen et al. (2017, 2018, 2020) and vibrations Coello Pérez and Papenbrock (2015a, 2016). Venerable nuclear collective models Eisenberg and Greiner (1970a); Bohr and Mottelson (1975); Iachello and Arima (1987) have been identified as leading-order Hamiltonians in an EFT approach.
In this work, we develop an EFT for odd-mass deformed nuclei. These are viewed as a nucleon coupled to an axially symmetric core. Many even-even deformed nuclei exhibit some amount of triaxiality even in low-lying rotational bands. That, however, is often a small effect that can be treated as a higher-order correction to a first-order description that uses axial symmetry. Our approach differs from the general particle-rotor model and from a very recently developed EFT Chen et al. (2020), both of which couple the nucleon to a triaxially deformed nucleus. As we will see below, the coupling of a nucleon to an axially symmetric core can, however, yield a triaxially deformed nucleus.
The theoretical arguments that lead to the Hamiltonian of the particle-rotor model are deceptively simple Herzberg (1945); Bohr (1952); Bohr and Mottelson (1953); Nilsson (1955); Kerman (1956): In the body-fixed (i.e. co-rotating) coordinate system (indicated here and in what follows by primes), a particle with angular momentum is coupled to a rotor with angular momentum , resulting in the total angular momentum . The Hamiltonian of a rotor is given by
(1) |
Here, are the moments of inertia. Replacing the components of by those of leads to the Hamiltonian
(2) |
of the particle-rotor model. That model describes a wealth of data on odd-mass nuclei.
We are motivated to develop an EFT for the particle-rotor model because that approach is expected to yield a systematic classification of terms in the Hamiltonian according to their order of importance, with the Hamiltonian (2) expected to appear as the leading-order term. For that, the formulation of the particle-rotor model in terms of angular momenta is not a good starting point, however. In line with common usage, our EFT is based upon the Lagrangian or the Hamiltonian formalism. These, in turn, make use of velocities or canonical momenta, respectively. However, a Lagrangian approach to the particle-rotor system is not contained in the standard textbooks Rowe (2010); Eisenberg and Greiner (1970a); Bohr and Mottelson (1975); Ring and Schuck (1980); Iachello and Arima (1987); Rowe and Wood (2010).
In addition to providing a systematic procedure for generating Hamiltonian terms of given order, the EFT approach yields surprises and interesting results. For example, the coupling between the particle and the rotor can naturally be described in terms of Abelian and non-Abelian gauge potentials. Such potentials, and the Berry phases Simon (1983); Berry (1984) associated with them, enter in the description of diatomic molecules Wilczek and Zee (1984); Jackiw (1986); Bohm et al. (1992) and the quantum Hall effect Estienne et al. (2011). However, Berry phases have received less attention in low-energy nuclear physics Nikam and Ring (1987); Bulgac (1990); Klein and Walet (1993); Nazarewicz (1994); Hayashi (1997); Chandrasekharan et al. (2008).
This paper is organized as follows. We identify the relevant low-energy degrees of freedom in Sect. II. In Sect. III we systematically construct the EFT by presenting the power-counting procedure and introducing the relevant interactions at leading and at next-to-leading order. Hamiltonian and total angular momentum are introduced, the Hamiltonian is diagonalized and spetra in leading and subleading order are calculated in Section IV. We present applications of our results to 239Pu and the triaxially deformed 187Os in Sec. V, and summarize our results in Sect. VI. Numerous appendices give the technical details necessary for a self-contained description.
II Degrees of freedom and separation of scales
II.1 Even-even nucleus: rotating core
Many odd-mass deformed nuclei can be viewed as an even-even deformed nucleus to which the extra nucleon is coupled. We take 239Pu as an example. The corresponding even-even nucleus is 238Pu, and Fig. 1 shows all its levels below 800 keV. At sufficiently low energies the spectrum of 238Pu is essentially that of an axially symmetric rigid rotor: The excitation energies versus angular momentum obey . Here is a rotational constant of about 7 keV, and keV (the energy of the state) sets the low-energy scale. Only even spins enter because the ground state is invariant under rotations by around any axis that is perpendicular to the symmetry axis. This symmetry is usually denoted as symmetry Bohr and Mottelson (1975). At energy keV a second rotational band with a band head occurs, followed by more rotational bands at higher energies. In this work we will, however, consider only the lowest energies and restrict ourselves to the description of the ground-state rotational band. Then, the energy of the band head sets the breakdown scale of our EFT, because a new degree of freedom enters at this energy. We have a separation of scale . The analysis of Ref. Coello Pérez and Papenbrock (2015b) shows that the ground-state band will exhibit noticeable deviations from the leading-order rule for spins . In the EFT this is due to subleading interactions that couple the ground-state band to other bands. While the interaction between the positive-parity ground-state band and the shown negative-parity band is suppressed, a positive-parity band enters at about 940 keV. These arguments suggest that the breakdown scale is properly chosen. An EFT for the lowest energies in deformed nuclei was presented in Refs. Papenbrock (2011); Coello Pérez and Papenbrock (2015b), and we briefly review its essential features.

II.1.1 The rotor in quantum mechanics
Nuclear deformation causes an emergent breaking Yannouleas and Landman (2007) of rotational symmetry from SO(3) to axial SO(2), described as a nonlinear realization of the symmetry Weinberg (1968); Coleman et al. (1969); Callan et al. (1969); Leutwyler (1994); Weinberg (1996a); Brauner (2010). The degrees of freedom corresponding to the remnants of Nambu-Goldstone bosons parametrize the coset SO(3)/SO(2), i.e. the two-sphere. We use the radial unit vector
(3) |
for this purpose. Here, are orthogonal unit vectors that span a right-handed coordinate system (the “space-fixed system”), and and are the polar and azimuthal angle, respectively. The vector in Eq. (3) points in the direction of the symmetry axis of the deformed nucleus. It is supplemented by the unit vectors
(4) |
The vectors span the (right-handed) “body-fixed” coordinate system of the rotor. They result from rotating the axes , , and of the space-fixed system by the operator . Here stands for the general rotation
(5) |
parametrized in terms of the Euler angles . The operators with generate rotations around the axes and fulfill the usual commutation relations
(6) |
We also use the notation
(7) |
for the basis vectors of the body-fixed coordinate system.
In addition to the generators of rotations in the space-fixed system we also use their analogues in the body-fixed system. These also obey the commutation relations (6). If space-fixed and body-fixed system originally coincide, the rotation (5) and the rotation
(8) |
are identical Varshalovich et al. (1988). For , the last two factors in expression (8) rotate the space-fixed -axis into the direction of . The remaining factor rotates the resulting system about the body-fixed -axis. Hence, an operator defined in the body-fixed system that is invariant under SO(2) rotations, is automatically invariant under general SO(3) rotations in the space-fixed system. We use that insight to construct invariant terms in the Lagrangian.
Our definition (II.1.1) of the body-fixed coordinate system, resulting from the application of the rotation to the space-fixed system, represents but one possibility. Any rotation with of the space-fixed system would be equally acceptable (albeit seems particularly simple). As we will see below, this arbitrary convention leads to a gauge freedom Littlejohn and Reinsch (1997).
The time-dependent angles describe the motion of the deformed nucleus. The angular velocity is
(9) | |||||
with
(10) |
The dot denotes the time derivative. We see that the rotor’s degrees of freedom transform non-linearly [i.e. they depend in a nonlinear way on ] under the rotation. The expression with defined in Eq. (9) is obviously invariant and so is, therefore, the Lagrangian
(11) |
This is, of course, the Lagrangian of an axially symmetric rotor (or, equivalently, that of a particle on the unit sphere). Here is a low-energy constant and corresponds to the moment of inertia.
We introduce the canonical momenta and and perform a Legendre transform of the Lagrangian (11). This yields the Hamiltonian
(12) |
Here, we combined the canonical momenta into
(13) |
We quantize the momentum as usual,
(14) |
with
(15) |
The spectrum is
(16) |
with angular momenta , corresponding to a rotational band.
An alternative derivation of the rotor spectrum uses the angular momentum
(17) |
rewrites, (which implies ), and thereby obtains the Hamilitonian . We will use such an approach below.
II.1.2 Connection to effective field theory
The arguments in Section II.A.1 may seem purely phenomenological. We now establish the connection to EFT. For nonrelativistic quantum systems, that approach is summarized in Ref. Brauner (2010), see also Ref. Leutwyler (1994). A paradigmatic application is that to the infinitely extended ferromagnet (FM) (Refs. Román and Soto (1999); Hofmann (1999); Bär et al. (2004); Kämpfer et al. (2005)). The breaking of a symmetry of the Hamiltonian in the ground state of the system (in the FM: the common direction of all spins violates rotational invariance) causes the existence of Nambu-Goldstone modes (in the FM: spin waves). These make up for the fact that the FM cannot rotate. They determine the low-lying part of the spectrum of the FM, are determined entirely by the broken symmetry, and depend upon a small number of parameters that must be fitted to the data. In atomic nuclei, that EFT scheme must be generalized as we deal with “emergent symmetry breaking”, see Refs. Yannouleas and Landman (2007); Papenbrock and Weidenmüller (2014); Papenbrock and WeidenmÃŒller (2015). In the limit of infinte mass, nuclei cannot rotate either. The Nambu-Goldstone modes are surface vibrations. Finite nuclei are able to undergo rotations, however. The associated degrees of freedom are the purely time-dependent angles and introduced in Section II.A.1. These degrees of freedom are not Nambu-Goldstone modes as they cease to carry physical significance in the infinite-mass limit. Rather they represent the onset of symmetry breaking in the finite system (hence “emergent symmetry breaking”). That approach is expected to work for systems close to full symmetry breaking. Then the relevant energy scales are (in increasing order) the rotational energy (via the degrees of freedom and ), the surface vibrations (described in terms of Nambu-Goldstone modes), and genuine intrinsic excitations of the system (not accessible in terms of Nambu-Goldstone modes). In Refs. Papenbrock and Weidenmüller (2014); Papenbrock and WeidenmÃŒller (2015) that approach has been worked out in detail for even-mass nuclei. In the present paper we confine ourselves in the description of the core to the very lowest part of the excitation spectrum, i.e., to rotations.
Needless to say we may re-formulate the quantum mechanics of the axially symmetric rotor as a quantum field theory. Based on the familiar rotor Hamiltonian (12) we introduce the quantum field operator that fulfills the canonical commutation relations for bosons
(18) |
Here, creates an axially symmetric rotor whose symmetry axis points into the direction of .
The Lagrangian of the free rotor is then
(19) |
Introducing the momentum operator
(20) |
and performing the usual Legendre transformation then yields the Hamiltonian
(21) |
This clearly is the second-quantized version of the Hamiltonian (12).
Although we are dealing with the rotor in quantum mechanics and not in quantum field theory we continue to use the terminology of EFT. This is in keeping with many works in low-energy nuclear physics where the ideas of EFTs Lepage (1997) are used to construct and solve Hamiltonians in quantum mechanics, see, e.g., Refs. Scaldeferri et al. (1997); Beane et al. (1998); Kirscher et al. (2010); Lensky et al. (2016); Capel et al. (2018).
II.2 Nucleon
To gain insight into how to construct the EFT, we consider the odd-mass nucleus 239Pu. Figure 2 shows all levels below 800 keV that can be grouped into rotational bands (omitting the few exceptions). The ground-state rotational band is built on a state, i.e., a neutron coupled to the 238Pu ground state. Rotations of this nucleon-nucleus state then produce the rotational band on top of the ground state. The first excited neutron state yields the state at keV, and its rotations produce the corresponding rotational bands. Thus, the fermion single-particle excitation energy is about half the breakdown scale in this nucleus, and the condition is fulfilled only marginally. The band head at about 470 keV could be due either to a single-neutron excitation or to the coupling of the neutron with the excited state (at the breakdown energy ) in 238Pu. Therefore, that rotational band is beyond the breakdown scale of the EFT we present in this paper.

The rotational bands depicted in Fig. 2 all follow the pattern
(22) |
Here, is an energy offset, the rotational constant, and the decoupling parameter (that occurs only for bands). These constants depend on the band under consideration. Typically, we have , , and . Equation (22) is well known from a variety of models Rowe (2010); Eisenberg and Greiner (1970b); Bohr and Mottelson (1975); Iachello and Arima (1987); Rowe and Wood (2010). As shown below, it is also the leading-order result of the EFT we develop in this paper.
We use the insight gained in the previous Subsection and request that the Lagrangian of the nucleon be invariant under SO(2) rotations in the body-fixed system. That guarantees invariance under SO(3) rotations in the space-fixed system.
The field operator creates a fermion at position with spin projection onto the -axis in the body-fixed frame. Denoting the vacuum as we thus have
(23) |
Here denotes a spin state of spin- fermion with projection Varshalovich et al. (1988), and is an eigenstate of the position operator. The corresponding annihilation operator is and we have the usual anti-commutation relation for fermions
(24) |
and all other anti commutators vanish. It will be useful to combine the two spin components of the field operator into the spinor
(27) |
The nucleon Lagrangian is
(28) |
Here, is an axially symmetric potential which may also depend on spin, i.e., be a matrix. The potential of the Nilsson model Nilsson (1955) is an example. The Lagrangian (28) exhibits axial symmetry. The construction (28) is not only mandated by the nonlinear realization of rotational symmetry Papenbrock (2011). It is also consistent with an adiabatic approach where the light nucleon is much faster than the heavy and slowly rotating core and able to follow the core’s motion quasi instantaneously. The canonical momentum is
(29) |
The Legendre transform of the Lagrangian (28) yields the Hamiltonian
(30) |
Here, we introduced the Laplacian . The total angular momentum of the fermion,
(31) |
is the sum of orbital angular momentum and spin
(35) |
Here, denote the usual Pauli matrices. These act with respect to the axes of the body-fixed system. The action of the general operators (, , on the space- and spin-degrees of freedom of the nucleon coincides with that of the corresponding angular momentum plus spin operators in Eq. (31). Thus, for ,
(36) |
are the projections of the fermion’s angular momentum onto the body-fixed axes.
Fermion states of axially-symmetric Hamiltonians are written as . Here denotes the projection of the angular momentum onto the nuclear symmetry axis, while comprises the remaining quantum numbers energy, parity, and third component of isospin. Kramers’ degeneracy (i.e., time-reversal invariance) implies that the single-Fermion states come in degenerate pairs , carrying identical quantum numbers and, in particular, sharing the same energy . Thus, we have
(37) |
We understand the band heads in Fig. 2 simply as energy eigenvalues of the fermion Hamiltonian with a suitably chosen potential . The symmetry of the nuclear ground state ensures that only a suitable linear combination of the states enters. The energies of the band heads are of the scale . Thus, if we had spontaneous rather than emergent symmetry breaking, would hold, and the rotational bands on top of each band head would collapse.
The degrees of freedom of the rotor do not appear explicitly in Eq. (30). Conversely, the potential has no impact on the degrees of freedom of the rotor. The potential constitutes an implicit interaction between the rotor and the particle which is solely based on the fact that the potential is axially symmetric and defined in the body-fixed frame. That is consistent with emergent symmetry breaking which allows only a coupling to derivatives of Nambu-Goldstone bosons, in our case: the angular velocity. Such interactions – not yet contained in the Hamiltonian (30) – will appear as gauge couplings of the nucleon to the rotor. These are partly constrained by the nonlinear realization of rotational symmetry. They appear in universal form as a covariant derivative or as Coriolis terms. They can also be understood within an adiabatic approach that involves Berry phases.
III Building an effective field theory
Having identified in Section II the relevant degrees of freedom due to a separation of scales, we now construct our effective field theory for odd-mass deformed nuclei using the following steps. (i) In the present Section we define the power-counting procedure and (ii) write down the interaction terms between the nucleon and the rotor in leading and some also in subleading order. In step (ii) all possible interaction terms are admitted that are allowed by the symmetries (in our case invariance under rotations, parity, and time reversal), see, e.g., Ref. Brauner (2010). (iii) The resulting Lagrangian and Hamiltonian, further subleading terms, and the solution of the equations of motion are addressed in Section IV.
III.1 Power counting
The power-counting procedure for the rotor was worked out in Refs. Papenbrock (2011); Papenbrock and Weidenmüller (2014). We briefly present the arguments. We associate the low-energy scale with the rotor. Thus, the angular velocity scales as ,
(38) |
and so does the Lagrangian (11) of the free rotor. That implies that its low-energy constant scales as
(39) |
The spectrum of the free axially symmetric rotor forms a rotational band, see Eq. (16), and is the moment of inertia. Let us give examples. MeV for a light rotor such as 8Be, 0.5 MeV in 24Mg, 0.2 MeV in the neutron-rich nucleus 34Mg, 30 keV for a rare earth nucleus, and only 15 keV for actinides. These are the smallest energy scales in the nuclei we consider. The breakdown energy for the rotor is set by excitations that are not part of its ground-state rotational band. This energy is about 17 MeV in 8Be, 4 MeV in 24Mg, 1 MeV in rare earth nuclei, and about 0.5 MeV in actinides. Thus, in all cases.
The subleading correction to the rotor Lagrangian (11) is
(40) |
At the breakdown scale, i.e., when , the term (40) is by definition equal in importance to the leading-order Lagrangian (11). That yields
(41) |
At low energy where , the term (40) yields a contribution to the total Lagrangian, and this is suppressed by compared to the leading term (11). That argument establishes the power-counting procedure for the rotor: the energy scale is associated with rotational bands. Corrections to the leading-order term come in powers of .
We turn to the energy scales of the Hamiltonian (30) of the nucleon. The energy scale is set by the mean level spacing of the single-nucleon states, i.e., by the spacing of band-head energies in odd-mass nuclei, see Eq. (II.2). That scale is about 1.7 MeV in 9Be, (given by the energy difference of the ground state and the excited band head), 0.6 MeV in 25Mg, (given by the energy difference between the ground state and the excited band head), and amounts to some hundreds of keV in rare earth nuclei, and to tens to hundreds of keV for actinides. In most odd-mass deformed nuclei we have , and in many cases one even finds , see 239Pu in Fig. 2 as an example. In such cases is only about a factor two or three away from the breakdown scale , and the separation between and is marginal so that but not . In such cases, the power counting uses both small expansion parameters and , and it is difficult to decide which is the more important one. If we had the ambition to construct an EFT for the nucleon potential in the Hamiltonian (30) we would have to deal, in addition, with an expansion in powers of , but we do not pursue this task in the present paper. In some nuclei such as 187Os discussed below, we have so that . Then, the power counting is in . [We avoid the equivalent parameter as that might incorrectly suggest a systematic expansion of the nucleon potential in the Hamiltonian (30).]
In any case, the separation of scales allows us to construct an EFT that systematicaly improves the energies and states in rotational bands. We note that there are many states in odd-mass nuclei that do not result from coupling a nucleon to the ground-state band of the even-even nucleus (but rather from coupling to excited band heads of the even-even nucleus). Such states fall outside the purview of the EFT we aim to construct. Including such effects would require us to introduce fields that describe the non-rotational excitations of the rotor.
It would be desirable to construct the potential in the fermion Hamiltonian (30) in a similarly systematic fashion. We briefly illuminate the difficulties in doing so for halo rotors, i.e., odd-mass nuclei where the nucleon is weakly bound to the even-even core. Examples are 9Be (with a neutron separation energy of about 1.7 MeV) and neutron-rich magnesium isotopes with separation energies below 1 MeV. In these cases, the fermion’s de-Broglie wave length exceeds the rotor’s size, and a derivative expansion of the potential seems appropriate. The potential must be axially symmetric. Total spin and its projection are trivial constants, while is the nontrivial conserved quantity and can be used to classify the fermion’s wave functions. Thus, we can parameterize the potential as
(42) | |||||
Here, . In Eq. (42) we did not present all second-order derivatives, and higher-order derivatives are missing as well. If the fermion has quantum numbers , the leading-order contribution consists solely of the contact coupling. For or states (e.g. for the ground state and excited band head at about 2.8 MeV, respectively, in 9Be), second-order derivatives in the potential must enter. In the latter case, one also needs to employ a potential that breaks spherical symmetry down to axial symmetry, thus lifting the four-fold degeneracy of a orbital in the body-fixed frame. The considerable number of low-energy coefficients then requires that a significant amount of data is available. In practice, one would like to adjust to scattering data (and make predictions for spectra and transitions), but those are rare. Odd-mass neutron-rich isotopes of magnesium, for instance, are expected to have ground states. This would require us to carry the expansion (42) to even higher order, and the scarcity of data in rare isotopes would prohibit us to follow such an EFT approach. It is, therefore, more practical to assume that the Hamiltonian (30) is already in diagonal form, with low-energy eigenstates as given in Eq. (II.2) fitted to the data. In other words, we take the single-particle energies of the fermion from data. This is somewhat similar in spirit to halo EFT Hammer et al. (2017) where each state of the core is represented as a separate field and simply adjusted to data.
The resulting EFT involves – in leading order – terms of order and . Subleading corrections are suppressed by factors of (or provided that holds). This EFT does not provide us with an expansion in powers of , because we do not construct such an expansion of the potential .
III.2 Nucleon-rotor interactions
We deal with emergent symmetry breaking. Thus, the nucleon can couple to the rotor only derivatively, i.e., via the angular velocity . All terms allowed by the symmetries must be considered. At face value, the resulting velocity-dependent couplings are well known. They involve – in the body-fixed frame – Coriolis forces. However, the essential physical argument for the couplings is more subtle and profound. The coupling terms are gauge couplings that involve Berry phases (or geometrical phases). Such phases occur in many quantum systems Simon (1983); Berry (1984). While originally conceived for systems that undergo a time-dependent adiabatic motion, they may also occur where “fast” degrees of freedom have been removed or integrated out, and where one is only interested in the remaining “slow” degrees of freedom Kuratsuji and Iida (1985). A well-known example from molecular physics is the Born-Oppenheimer approximation. Here, Berry phases and the corresponding gauge potentials enter the dynamics of the nuclei of the molecule once the electronic degrees of freedom have been removed. That leads to the molecular Aharonov-Bohm effect Mead and Truhlar (1979); Wilczek and Zee (1984). For the diatomic molecule, some details are presented in Refs. Jackiw (1986); Bohm et al. (1992); Rho (1994). In the present case, the fact that the nucleon is much faster than the slowly rotating core shows that gauge potentials play a role. Likewise, gauge potentials are a general feature of systems where a separation between rotational and intrinsic degrees of freedom is being made, and different conventions for this separation differ by gauge transformations Littlejohn and Reinsch (1997).
The non-linear realization of rotational invariance requires that, in the body-fixed system, we have to replace the time derivative in the Lagrangian (28) by the covariant derivative Papenbrock (2011) (see Appendices C and D)
(43) |
Here, the commutator’s second argument is left open. The last term of the covariant derivative accounts for Coriolis effects in the body-fixed system. It is present even if the Lagrangian in the body-fixed system does not depend on time explicitly, i.e., even if the partial time derivative vanishes. In the Lagrangian (28) that yields
(44) |
We have factored out the angular velocity , see Eq. (9), and we have used Eq. (36). This naturally introduces the gauge potential
(45) |
which couples the rotor to the nucleon via . Here we borrow the expression gauge potential from electrodynamics. In the parlance of differential geometry, the field is a connection. The term scales as , i.e. the gauge potential is dimensionless and of order one. Thus, the gauge term is as important as the Lagrangian (11) of the free rotor and enters in leading order.
The gauge potential (45) is singular at the north and south poles of the unit sphere. Single-valuedness of the wave function for the rotor requires that the eigenvalues of be integer or half integer. That is obviously guaranteed for the fermion for which , , . We compute the corresponding magnetic field (or the Berry curvature in differential geometry) and find
(46) |
This is the field of a Dirac monopole on the unit sphere and clearly exhibits spherical symmetry Fierz (1944); Wu and Yang (1976), in contrast to the gauge potential (45) whose rotational invariance is not obvious. As shown in App. E.2, the effect of a rotation on the gauge potential (45) can be reversed by a gauge transformation.
To see that the field (46) is indeed a monopole field we take a detour and consider a sphere of radius of the size of the nucleus, with angular and radial coordinates and , respectively. We neglect excitations of the sphere in the radial direction as these relate to vibration with energies at the breakdown scale and put . Then angular velocities and the vector potential (45) are multiplied with and , respectively. The usual differential operator then shows that we deal indeed with a monopole field.
The gauge potential (45) is intimately linked to the geometry of the sphere, i.e. the coset space SO(3)/SO(2). To see this, we consider a sequence of three rotations (around space-fixed axes) that take the rotor from a point on the unit sphere to a point , then from to a point , and finally from point back to the point . We assume that the three rotations are around three distinct axes. This ensures that the triangle on the sphere has a finite solid angle (or area). It is clear that at least two of the three rotations will also induce rotations of the fermion field around the body-fixed -axis. While the rotor has returned to its original configuration after the sequence of the three rotations, the fermion’s configuration has been changed by a finite rotation around the body-fixed -axis. Inspection shows that the rotation angle, i.e. the phase acquired by the fermion with spin projection , is equal to the solid angle of the enclosed loop times the spin projection . Dynamically this phase is acquired because of the monopole magnetic field.
Another gauge coupling, permitted in the framework of our EFT, is
(47) |
Here, is a dimensionless coupling constant. Its natural size is of order unity, and the contribution (47) scales as and enters at leading order. The corresponding gauge potential is
(48) |
This vector potential contains non-commuting operators and, therefore, constitutes a non-Abelian gauge potential. The corresponding magnetic field is
(49) | |||||
Taken by itself, this magnetic field is not invariant under rotations. However, the total gauge potential is
(50) |
and the corresponding magnetic field
(51) | |||||
is spherically symmetric. Again, we deal with a magnetic monopole. However, in contrast to the field (46) its overall strength is not quantized because the non-Abelian vector potential (48) exhibits no singularities on the unit sphere and the coupling can therefore assume any real value.
To show how this gauge term relates to the Berry phase we observe that the states have the same energy, see Eq. (II.2). When the rotor moves along a closed loop on the unit sphere, a general interaction would mix the two degenerate states while transversing the loop. Thus, the initial and final fermion states could differ by a unitary transformation. The non-Abelian gauge potential (48) generates such a mixing for non-zero values of the coupling .
The choice of the gauge potentials is not unique. At any orientation of the rotor’s symmetry axis, the body-fixed coordinate system is defined up to an arbitrary rotation around the axis. Thus, the intrinsic degrees of freedom (of the fermion in our case) depend on a convention which is arbitrary. Different choices of the body-fixed system lead to different expressions for the covariant derivative and to different gauge potentials Littlejohn and Reinsch (1997). These are related to each other by gauge transformations. Details are given in Appendix E.
IV Lagrangian and Hamiltonian
IV.1 Leading- order terms
Collecting the leading-order results from the previous Sections, we find that the Lagrangian is given by
(52) | |||||
Here, is defined in Eq. (28). The Legendre transform of the Lagrangian (52) yields the Hamiltonian. For that was done in Section II.2. For the remaining variables the transformation is less tedious than might appear at first sight. The Lagrangian (52) is a quadratic form in the velocities and can be written as
(53) |
where T denotes the transpose. That Lagrangian has the Legendre transform
(54) |
Here, is a “mass” matrix and denotes its inverse. In the present case the canonical momenta are
(55) |
Here, we employed Eq. (II.1.1). The Hamiltonian becomes
(56) | |||||
Here, the fermion Hamiltonian is given in Eq. (30). The momentum is conserved because does not appear in the Hamiltonian (56). Combining two of the canonical momenta (IV.1) into
(57) |
we find
(58) |
Using that, we write the Hamiltonian (56) in compact form as
IV.2 Angular momentum
Replacing the canonical momenta by the total angular momentum simplifies the Hamiltonian and establishes the connection to Eq. (2). In the present Subsection we introduce the total angular momentum on an intuitive basis. A derivation based on Noether’s theorem is given in App. B.3.
The total angular momentum
(60) |
is the sum of the angular momentum of the fermion,
(61) |
which points in the direction of the symmetry axis, and that of the rotor,
(62) | |||||
which is perpendicular to it. In the last line of Eq. (62) we have used Eq. (58), see also Refs. Fierz (1944); Wu and Yang (1976). The term is the angular momentum of the rotor. The gauge potential is not manifestly invariant under rotations but can be made so via a gauge transformation Fierz (1944). That causes the correction (61) in the direction of . The equality
(63) |
shows that the conserved momentum is the usual angular momentum with respect to the space-fixed axis. We use Eqs. (62) and (60) to express the angular velocity in terms of the angular momentum. That yields
(64) |
Using that in the rotational energy , we arrive at the Hamiltonian
(65) | |||||
The term proportional to in Eq. (65) might be absorbed into (and then be dropped). The rotational part displayed in the second line of Eq. (65) corresponds to the rotor model (2) for the special case of axial symmetry, i.e. for .
IV.3 Spectrum
Simplifying the notation used in Eq. (II.2) we denote the ground state of the fermionic part of the Hamiltonian (65) as . We calculate the eigenfunctions of the rotor part of the Hamiltionian (65) by determining the eigenfunctions of , , and . For states we have
(67) |
The quantization proceeds as in Section II. The eigenfunctions of are
(68) |
The negative eigenvalue is chosen here to be consistent with chapter 4.2 of Ref. Varshalovich et al. (1988). The eigenfunctions of the square of the total angular momentum operator can be written either in terms of Wigner functions or in terms of Wigner functions (see chapter 4 of Ref. Varshalovich et al. (1988)). These are related by
(69) |
For we have
(70) | |||
(71) |
For the Hamiltonian (65) that implies
(72) |
Discrete symmetries of the rotor-plus-fermion system may select a definite linear combination of the states . These share the absolute value and have the same energy [see Eq. (II.2)]. Combining with Eq. (IV.3) yields
(73) |
The term linear in of the Hamiltonian (65) couples states and . For most heavy nuclei where , the coupling of states with different values of is of subleading order and can be computed perturbatively. For , however, the interaction couples the degenerate states and is, thus, of leading order . For this case, eigenfunctions and eigenvalues are worked out in Appendix F. The result for the eigenvalues,
(74) | |||||
agrees with Eq. (22) when we express the constants , , and in terms of , , and . The last term in Eq. (74) is known as the signature splitting. Equation (74) shows that for the spectrum changes by about . For , the spectrum would resemble a rotational band only for . This confirms that the natural size of is of order unity.
States that differ by one unit in can also be coupled strongly by the term linear in in the Hamiltonian (65) provided . In that case, the spectrum can be calculated analytically using as a basis the eigenstates obtained for and taking account only of the two bandheads. The diagonalization of the matrix spanned by the states and yields the eigenvalues Kerman (1956)
The energies are given by Eq. (74), and is a low-energy constant. The sign on the right-hand side of Eq. (IV.3) has to be chosen such that the energies and for the bands with quantum numbers and , respectively, are obtained as . In nuclei such as 105,107Mo, groups of more than two band heads are closely spaced and strongly coupled. In such cases, a Hamiltonian matrix of larger dimension needs to be diagonalized.
We discuss our results. For , the total angular momentum and its projections and onto the space- and body-fixed -axes, respectively, commute with each other and with the Hamiltonian. The spectrum is given by Eq. (73). The nucleus is axially symmetric because is conserved. For finite , the projection of the angular momentum onto the rotor’s symmetry axis is not conserved because the Abelian and non-Abelian gauge potentials do not commute. According to the rules for power counting, the term linear in (the “Coriolis term”) is of leading order. Nevertheless, the impact of the Coriolis term on the spectrum depends very much on the nucleus under consideration. In a band with band-head spin this term contributes of the order . Thus, it is only of leading order for a rotational band with . However, the Coriolis term also couples bands that differ in by one unit. Equation (IV.3) shows that the Coriolis term is of leading order only if is sufficiently large, i.e. of order unity. In practice, this is mostly expected if two band heads that differ in spin by one unit are closely spaced in energy. Here “close” means that the spacing is not of the typical fermion scale but rather of the rotational scale . In the presence of the Coriolis term, is not a conserved quantity anymore, and the odd-mass nucleus exhibits triaxial deformation. We illustrate this behavior below for 187Os. From the point of view of our EFT, triaxiality in odd-mass nuclei thus depends on the spins of band heads and on their separation in energy.
IV.4 Next-to-leading order corrections
The leading-order Hamiltonian (65) contains contributions that scale as and/or . Out of the many terms quadratic in both and that one can write down using , , and , the following combinations are linearly independent and compatible with the symmetries:
(76) |
The natural assumption is that . Then, the contributions of scale as , which is a factor smaller than the leading-order Lagrangian (52). The next-to-leading order terms (IV.4) are still quadratic in the velocities. After adding these terms to the Lagrangian (52) we can, therefore, perform the Legendre transform as outlined in Subsection IV.1, but invert the mass matrix perturbatively. The calculation is done in Appendix F. The resulting Hamiltonian is
(77) |
with the leading-order Hamiltonian as in Eq. (65) and with
(78) |
The dimensionless operators and are all of order and depend on bilinear combinations of the fermion operators , , and . In Eq. (78) we also used
(83) |
The matrix , due to , is diagonal in the eigenstates of the leading-order Hamiltonian (65). Thus, the first term on the right-hand side of Eq. (78) adds a fermion-state dependent correction of order to the moment of inertia. It causes the moments of inertia of rotational bands in odd mass nuclei to deviate somewhat from the moment of inertia for the ground-state band of the even-even rotor. The correction can be compared to the smaller variations of order that occur in even-even nuclei Zhang and Papenbrock (2013). The matrix (due to ) in the second term is traceless and mixes fermion states that differ in quantum numbers by two units. In particular, this term modifies the rotational spectra of band heads.
How does our approach compare with a treatment that would use full-fledged quantum field theory? While in the derivation of the EFT we employed velocities and canonical momenta, the solution of the Hamiltonian became simple because we introduced angular momenta. In the gauge we used the eigenfunctions are the Wigner functions (69). These can be written as infinite sums of spherical harmonics, i.e. of the “free” solutions of the even-even rotor. We are convinced that using , gauging the quantum-field theory Lagrangian (19 with the gauge potential (50), and using field-theoretical tools such as Feynman diagrams, would yield the same result. Then, the “free” rotor would scatter via infinite loops with the fermion, with vertices due to the gauge coupling.
V Applications
In the previous Section we have shown that in leading order, the EFT for odd-mass deformed nuclei recovers the results of the (axially symmetric) particle-rotor model. While that model is well known, with numerous applications to be found in textbooks Rowe (2010); Eisenberg and Greiner (1970b); Bohr and Mottelson (1975); Iachello and Arima (1987); Rowe and Wood (2010) and in the literature, the EFT provides us, in addition, with a systematic approach to subleading corrections and to estimates of the uncertainty of EFT predictions Furnstahl et al. (2015). We illustrate that point, following arguments made previously for even-even deformed nuclei with axial symmetry Coello Pérez and Papenbrock (2015b) and for vibrational exitations in heavy nuclei Coello Pérez and Papenbrock (2015a, 2016).
V.1 239Pu
Within the EFT the nucleus 239Pu is described as a neutron attached to 238Pu. Inspection of the low-lying states of 238Pu in Fig. 1 shows that the low-energy scale is keV and the breakdown scale is keV. This is probably too conservative an estimate for the breakdown scale of the ground-state band in 238Pu, because the lowest band head with positive parity occurs at 941 keV. Thus, for a description of the ground-state band, is probably a more accurate estimate for the the expansion parameter. Adjusting the low-energy constant to the energy of the state yields keV. The leading-order EFT predictions Papenbrock (2011); Coello Pérez and Papenbrock (2015b) for the ground-state rotational band are levels at energies
(84) |
Here, we included the EFT uncertainty estimate Papenbrock (2011); Coello Pérez and Papenbrock (2015b). Figure 3 compares the EFT results to data. For the uncertainty estimate we used , where the factor 0.25 is determined empirically.

In leading order, the rotational constant of the nucleus 239Pu is the same as for 238Pu. We only have to adjust the constant in Eq. (74) to describe the ground-state band. A fit to the first excited state in this nucleus yields . The resulting ground-state band is shown in the left part of Fig. 4 and compared to data in the center. At leading order, the rotational constant has a relative uncertainty of , as reflected by the blue shaded areas. For the displayed uncertainties, we used the conservative estimate and , with the factor of 2 determined empirically.
A next-to-leading order fit to the energies of Eq. (74) is shown in the right part of Fig. 4. Here, we adjusted both and in Eq. (74), finding keV and . We note that the change of by about a factor of is consistent with EFT expectations. At next-to-leading order, relative energy uncertainties are estimated as with . As before, the factor 0.25 is determined empirically.

We see that the EFT yields an accurate (it agrees with the data within the uncertainties) and increasingly precise (as more orders are included) description of the ground-state rotational band of 239Pu. Furthermore, the low-energy constants are not merely fit parameters, but the size of subleading corrections can be estimated from the empirical values of the low-energy scale and the breakdown scale . Similar results can also be obtained for the other rotational bands displayed in Fig. 2.
V.2 187Os

In most odd-mass nuclei, the Coriolis term [last term in Eq. (65)] that couples different rotational bands enters only perturbatively, because band heads that differ in by one unit are usually an energy apart. However, in nuclei with closely spaced band heads, the Coriolis term is of leading order. Among these are the light nucleus 9Be, the nuclei 49Cr and 49Mn, 105,107Mo, 187Ir, and 187Os. We illustrate our results for the well-studied nucleus 187Os Malmskog et al. (1971); Morgen et al. (1973); Sodan et al. (1975). The ground state exhibits a rotational band with a low-energy constant keV. The first excited band head is only separated by keV. Thus, we have , the two bands in question are coupled by the Coriolis term, and Eqs. (IV.3) must be employed.
The relevant scales are as follows. The even-even nucleus 186Os exhibits a ground-state rotational band with a state at 137 keV; the excited band head at 770 keV sets the breakdown scale of this rotor. The ratio of the energies of the two lowest states is . In a first step we neglect the coupling between the and bands in 187Os. Adjusting a total of five parameters [, , and in Eq. (74)] to the lowest three states for and fitting separately and to the two states of the band yields the two rotational bands shown in left part of Fig. 5. Here, the highest two (three) states of the () band, respectively, are predictions. The results are to be compared to the data shown in the center. Also shown in the right part are the EFT predictions obtained by adjusting the five parameters , , , and in Eq. (IV.3) simultaneously to the lowest three states of both bands. Given the same number (five) of low-energy constants, the improved accuracy obtained in the second fit shows the need to include the Coriolis coupling. Comparing the results to the data we infer that relative EFT uncertainties are about %. Figure 6 shows the energy differences beween theory and data for both bands using EFT (blue) and neglecting the coupling between the bands (black). We see that approach that neglects the coupling between the bands loses accuracy as soon as one considers states that were not fitted.

VI Summary
We have developed an effective field theory for deformed odd-mass nuclei. In this approach, the odd nucleon experiences an axially-symmetric potential in the body-fixed frame of the even-even deformed nucleus (a rotor). The power counting is based on the separation of scales between low-lying rotational degrees of freedom on the one hand and both, higher-lying nucleonic excitations and intrinsic excitations of the even-even nucleus, on the other. In leading order, the nucleon is coupled to the rotor via gauge potentials. Actually, the non-Abelian gauge potential is a truly first-order term only for band heads or when band heads with quantum numbers that differ by one unit of angular momentum, are close in energy. In the latter case, the gauge potential induces triaxiality. That was shown by applying the EFT to 187Os. We have shown how subleading contributions can be constructed systematically, and how these may be used to improve the spectrum and/or to estimate theoretical uncertainties. The EFT developed in this paper presents a model-independent approach to the particle-rotor system that is capable of systematic improvement.
Acknowledgements.
This work has been supported by the U.S. Department of Energy under grant No. DE-FG02-96ER40963 and under contract DE-AC05-00OR22725 with UT-Battelle, LLC (Oak Ridge National Laboratory).Appendix A Overview
Appendix B presents details regarding transformation properties under rotations. In App. C we derive the expression for the covariant derivative. Appendix D presents a more formal derivation of these properties based on the coset approach. In App. E we discuss gauge potentials and gauge transformations. App. F presents details regarding the derivation of the spectrum and subleading corrections.
Appendix B Transformation properties under rotations
In this Appendix we use infinitesimal rotations and apply Noether’s theorem to derive the expressions in Section IV for the total angular momentum, both in the space-fixed and in the body-fixed system.
An infinitesimal rotation changes the angles to . That moves the symmetry axis of the rotor into a new direction, and it induces a rotation of the axes and around the rotors new symmetry axis by an angle . In the following two subsections, we relate the infinitesimal angles to the parameters of a general infinitesimal rotation. We do so for rotations around the axes of the space-fixed system and for rotations around the body-fixed axes. We use that at the point , the body-fixed basis vectors are
(85) | |||||
B.1 Rotations around the space-fixed axes
A rotation by the vector about infinitesimal angles , around the space-fixed axes changes the body-fixed basis vectors , by
(86) | |||||
We equate the incremental changes of on the right-hand side of the last of Eqs. (B) with the last line of Eq. (B.1). That yields
(94) |
The rotor’s degrees of freedom clearly transform non-linearly, i.e., under the rotation by they depend in a nonlinear way on . The rotated basis vector differs from the basis vector by a small rotation with the angle around the rotor’s symmetry axis . To determine we compute the scalar product
(96) |
B.2 Rotation around the body-fixed axes
A rotation by the vector about infinitesimal angles , around the body-fixed axes changes the body-fixed basis vectors by
(111) | |||||
Equating the incremental change of on the right-hand side of the last of Eqs. (B) with the last line of Eq. (B.2) gives
(119) |
The incremental rotation angle is given by the scalar product of the rotated basis vector and the basis vector ,
(120) |
That shows that a rotation by points the body-fixed system into the new direction and rotates the body fixed system around its new axis by the angle . The fermion wave function transforms as in Eqs. (97, 98) but with replaced by . Thus,
(127) |
with
(131) |
B.3 Noether’s theorem and angular momentum
We use Noether’s theorem Noether (1918) and the results of Subsections B.1 and B.2 to obtain expressions for the conserved quantities, i.e., the components of total angular momentum in the space-fixed and in the body-fixed system, respectively. These are used in Section IV of the main text.
The theorem expresses invariants of the system in terms of partial derivatives of the Lagrangian with respect to the velocities of the system. The Lagrangian is given by the second of Eqs. (52), with defined in Eq. (28). The velocities are , , and the time derivative of the fermion wave function. The conserved quantities are the components of angular momentum, expressed in terms of the transformation matrices of Eqs. (109) and (131) and given by
(133) |
for rotations around the space-fixed axes and
(134) |
for rotations around the body-fixed axes. In the case of the velocity , the summations on the right-hand sides of Eqs. (133) and (134) actually involve an integration over . We use Eq. (IV.1), perform the space integration over the matrix elements of , and use Eq. (36). In the space-fixed system we find
(135) |
In the body-fixed system we have
(136) |
The square of the total angular momentum, defined by the sum of the squares of its components and calculated either in the space-fixed or in the body-fixed system, in both cases is given by Eq. (66). Upon quantization, the components (B.3) do not fulfill the canonical commutation relations as they are not generators of rotations. For a discussion of unusual commutation relations we refer the reader to Ref. Nauts and Gatti (2010).
Appendix C Covariant derivative
For the time derivative of a vector in the tangential plane of the two-sphere at , we use
(137) |
and have
(138) | |||||
The projection of onto the tangential plane defines the covariant derivative
(139) | |||||
The covariant derivative consists of the usual time derivative and a rotation in the tangential plane, i.e., a rotation by around the axis.
It is straightforward to generalize this argument to spin functions. Let with spin and projection be the spin function in the space-fixed system. A rotation to the body-fixed system yields the helicity spin states
(140) |
These are quantized with respect to the body-fixed axis, see chapter 6.1.3 of Ref. Varshalovich et al. (1988). The time derivative is
(141) | |||||
Here we used formulas from chapter 4.9 of Ref. Varshalovich et al. (1988). We also find
This allows us to perform the sum, and we arrive at
(143) |
Here, we used . To obtain the part relevant for the covariant derivative we project the right-hand side of Eq. (143) back onto . For a general spin function in the body-fixed system we thus have
(144) |
Had we written the vector considered above in terms of its spherical components we would have obtained the same result. Applying the result (144) to the spinor functions yields Eq. (43).
Appendix D Coset space
We exploit the nonlinear realization of rotational invariance more formally than done in the calculations of Apps. B and C. Thereby we connect to previous EFTs on axially deformed nuclei Papenbrock (2011); Papenbrock and Weidenmüller (2014); Coello Pérez and Papenbrock (2015b), nuclei with tri-axial deformation Chen et al. (2017, 2018, 2020), and magnets Román and Soto (1999); Hofmann (1999); Bär et al. (2004); Kämpfer et al. (2005). We follow closely the original papers Weinberg (1968); Coleman et al. (1969); Callan et al. (1969). For reviews of this approach, and an exhibition for non-relativistic systems, we refer the readers to Refs. Leutwyler (1994, 1997); Brauner (2010) and the textbook Weinberg (1996b). In finite systems, one speaks of emergent symmetry breaking Yannouleas and Landman (2007) but the tools from field theory can also be extended to this case Gasser and Leutwyler (1988); Papenbrock (2011); Papenbrock and Weidenmüller (2014). Not surprisingly, the calculations in the present Section have much in common with those in Appendices B.1 and B.2.
Three mutually orthogonal unit vectors (the “body-fixed system”) are linked to another three mutually orthogonal unit vectors , , (the “space-fixed system”) by a rotation so that for we have . That can be written as where is the matrix representation of . The matrix is real orthogonal, , hence . Altogether,
(145) |
For vectors we use small (capital) letters when they are written in the space-fixed (the body-fixed) system, respectively. For a vector we have where
(146) |
The transformation is defined as
(147) |
It coincides with the transformation in Section II.1. The three generators of infinitesimal rotations about the -axes obey
(148) |
The matrix representation of the operators is
(149) |
The commutation relations (148) for the matrix representation are verified using standard matrix algebra. The relations (149) imply
(150) |
and, thus,
(151) |
In the body-fixed system, we define a set of three operators , . These have the same commutation relations (148) as the operators . Moreover, these operators have, by definition, the same matrix representation (149) in the basis as do the operators in the basis . Hence, with
(152) |
we have for
(153) |
The commutation relations for the operators differ in sign from the anomalous commutators commonly used in the body-fixed system. The reason is that the definition (153) employs the matrix representation of on its right-hand side. Conventionally, when using differential operators for , these act also on the angles in and one obtains additional transformation terms leading to anomalous commutation relations. The operators and differ. That is seen by comparing the matrix representations in the basis ,
(154) |
In analogy to Eq. (147) we define the operator
(155) |
In the body-fixed system, the matrix elements of are given by
(156) |
Eq. (153) implies that the matrix elements of in the space-fixed system and of in the body-fixed system are equal,
(157) |
The equality of these two matrices implies that we may use either form. If the matrix operates in the space-fixed system we use the form , if it acts in the body-fixed system, we use the form . If we employ an operator representation we proceed likewise and use as defined in Eq. (148) in the space-fixed system and as defined in Eq. (155) in the body-fixed system.
Let the angles and, with these, the transformation be dependent upon time. Let be a vector in the tangential plane (i.e., perpendicular to ) with time-dependent components . The time derivative of , indicated by a dot, is
(158) |
We use and . Moreover, from it follows that the matrix is antisymmetric. Thus,
(159) |
Explicit calculation shows that
(160) |
Combining Eqs. (153) to (155) we obtain
(161) | |||||
This is Eq. (138). The covariant derivative of is defined as the projection of onto the tangential plane,
(162) |
Using the fact that in the basis the operators have the matrix representation (149), we write Eq. (162) in the form
(163) |
That agrees with Eq. (43). The partial derivative acts only on the components of . The additional term accounts for a rotation around the -axis by the angle . That is the hallmark of a covariant derivative.
Given two vectors and in the space-fixed system with time-dependent coefficients , we transcribe the inner product of and of the time derivative of , i.e., the expression , into the body-fixed system. As mentioned earlier we distinguish the system-dependent representations of the two vectors by writing and . To focus attention on the covariant derivative we put . Then both and are tangential vectors in the body-fixed system. From Eq. (146) we have , and, thus,
or, using the definition (163),
(165) |
Eq. (165) gives the rule for transcribing time derivatives of vectors into the body-fixed system. It applies provided in the body-fixed system, the vectors are tangential.
We define an infinitesimal transformation in the space-fixed system and another infinitesimal transformation in the body-fixed system. Both are defined in terms of the augmented rotation . Here are infinitesimal. That changes . In the space-fixed system we consider the infinitesimal transformation acting on the vectors , keeping the vectors fixed. Eqs. (145) give
(166) | |||||
The last relation defines . Explicit calculation shows that
(167) | |||||
A general infinitesimal transformation in the space-fixed system can be written in terms of infinitesimal angles as
(168) |
Equating that with as given in Eq. (167) we obtain a linear relation between the infinitesimal angles , , and the angles . It reads
(178) | |||
(179) |
The inverse relation is
(189) | |||
(190) |
Identifying with , we see that this agrees with Eqs. (94) and (B.1). In the body-fixed system we consider the infinitesimal transformation acting on the vectors , keeping the vectors fixed. Equations (145) give
(191) | |||||
The last relation defines . Eqs. (191) show that acts in the body-fixed system. We use the arguments below Eq. (157) to express in terms of the operator defined in Eq. (155). Explicit calculation shows that
(192) | |||||
Since we have . The last relation shows that the three infinitesimal angles , , all carry negative signs. That is because the infinitesimal transformation acts conversely to the infinitesimal transformation . A general infinitesimal transformation in the body-fixed system can be written in terms of infinitesimal angles as
(193) |
Since in Eq. (192) acts conversely to we equate expression (192) not with expression (193) but with the converse of expression (193), obtained by the replacements for all . That gives
(203) |
The inverse relation is
(213) |
Identifying with , we see that this agrees with Eqs. (119) and (B.2).
The commutation relations (148) imply that under the infinitesimal rotation , the operator is mapped onto , and correspondingly for the other components. That shows that under a rotation, the three operators transform like the three unit vectors of a three-dimensional linear space. That suggests that in Eq. (168) and in Eq. (193) represent the same vector written, respectively, in the space-fixed and in the body-fixed coordinate system. For that to be true the three infinitesimal angles of and of must be connected as in Eq. (146),
(214) |
Combining Eqs. (179) and (213) yields
(218) | |||
(225) | |||
(226) |
Equation (151) shows that Eq. (226) indeed equals Eq. (214), confirming the vector character of . Applying that to Noether’s theorem in App. B.3 we see that and are indeed the components of the same vector written respectively, in the space-fixed and in the body-fixed coordinate system.
It is straightforward to extend these arguments from vectors, i.e. spherical tensors of rank three, to spherical tensors of arbitrary rank. Then, the concrete representations of the rotation matrices and are given by Wigner matrices, while all algebraic relationships derived above remain unchanged.
Appendix E Gauge potentials
We demonstrate how gauge potentials arise in an adiabatic approach and we discuss gauge transformations and their relation to rotations.
E.1 Gauge potentials from an adiabatic approach
The appearance of the non-Abelian gauge potential (45) can be understood also in an adiabatic approach Kuratsuji and Iida (1985); Wilczek and Shapere (1989). If the nucleon’s degrees of freedom are much faster than those of the rotor, the eigenstates of the fermion Hamiltonian follow the rotor’s axial symmetry instantaneously, independently of any details of the fermion-rotor interaction. For simplicity we consider only the fermion spin function with half-integer spin and projection onto the space-fixed axis. As the fermion is fast, it’s spin is in an eigenstate with respect to projection onto the rotor’s symmetry axis, i.e. the helicity spin states from Eq. (140) span, for fixed projection , a basis of the instantaneous fermion eigenstates. They fulfill
(227) |
Due to Kramers’ degeneracy, the spin states are degenerate. In the adiabatic approximation, one evaluates the Hamiltonian of the fermion-plus-rotor system
(228) |
in these eigenstates to get the effective Hamiltonian matrix (see, e.g., Berry’s overview in Ref. Wilczek and Shapere (1989))
(229) | |||||
Here the vector gauge potential is the matrix
(230) |
Using properties of Wigner functions Varshalovich et al. (1988) and a summation formula from Ref. Lai et al. (1998), one finds
(231) |
In our case, the projection is obtained by application of the operator , and we thus find that the gauge potential enters. This is Eq. (45).
E.2 Gauge transformations
Let us also explore gauge transformations. Our definition of the body-fixed coordinate system (II.1.1) is convenient [because the basis vectors and are tangent vectors of the lines parameterized by the spherical coordinates ] but otherwise arbitrary. Any rotation of these basis vectors around the axis would have been equally valid, i.e. the vectors
(232) |
and span a right-handed body-fixed coordinate system. Here, is a smooth function over the sphere. Let us repeat the computations made in the previous Subsection for this body-fixed system. The helicity basis functions for the fermion become
(233) |
Here, and in what follows we suppress the dependence of on the angles . The gauge potential is
(234) | |||||
We have used results from chapter 4.9 of Ref. Varshalovich et al. (1988). The sums over the last two terms cancel each other, and we used and from Ref. Lai et al. (1998).
The vector potentials and differ by a gauge transformation, which is generated by the arbitrary angle . For , for instance, one obtains the gauge potentials by Wu and Yang (1976). Another interesting choice is , because it generates the non-singular gauge potential . Our gauge potential (45) is singular at both poles, and the rotor eigenfunctions are Wigner functions. As pointed out in Ref. Dray (1986), the different gauge potentials correspond to different conventions regarding the third argument of the Wigner function, i.e. to different conventions regarding rotations of the body-fixed coordinate system around its axis. In other words, the wave function we used in the main text is replaced by when a gauge transformation is made.
The arguments of this Subsection show that the freedom of choice of the intrinsic coordinate system introduces a gauge freedom in the dynamics of the collective rotational degrees of freedom Littlejohn and Reinsch (1997). The general Abelian gauge potential is
(235) |
We extend the discussion to the non-Abelian gauge field. In its manifestly gauge-invariant form it reads
We have employed the operators for in the body-fixed frame. In the last line we have expressed the result in the coordinate system of the rotor’s angular velocity. We have used that the spin operators , , and fulfill the canonical commutation relations. Comparison with Eq. (48) meets our expectations, because the spin operators are simply rotated by around the body-fixed axis.
The magnetic field of the total gauge potential is
(237) | |||||
in agreement with Eq. (51). It does not depend on and is, thus, gauge invariant.
We turn this argument around and study the effect of rotations on the gauge potential. While the non-Abelian part (E.2) of the gauge potential is manifestly invariant under rotations, this is not so for the Abelian part (45). Under a rotation we have
Here, the differential is given in Eq. (B.1) for rotations around the space-fixed axes. This rotated gauge potential has to be compared with the gauge potential
(239) | |||||
at the point . Here, the differential is taken from Eq. (94). The difference
(240) | |||||
can be written as when employing the expressions (94) and (B.1). Thus, after a rotation the gauge potential can be brought back into its original form (50) by performing a gauge transformation Fierz (1944).
Appendix F Supplements to Section (IV)
To compute the contribution of the term linear in of the Hamiltonian (65) for states we introduce spherical components
(241) | |||||
and
(242) |
We write the term as
(243) |
Using the properties of the raising and lowering operators (see chapters 3.1 and 4.2 in Ref. Varshalovich et al. (1988)) we find
(244) |
and
Thus,
(246) |
Inspection shows that that the linear combinations
(247) |
are solutions of the Hamiltonian (65) for . The phase results from the requirement that the odd-mass nucleus is invariant under rotations by around any axis perpendicular to the symmetry axis. Hence, the contribution from the term proportional to in the Hamiltonian (65) becomes
(248) |
That yields Eq. (74).
We next compute the matrix elements of the -dependent terms of the Hamiltonian (65) for two close-lying band heads. Using the normalization to of the squared Wigner function, see Chapter 4.11 of Ref. Varshalovich et al. (1988), we find
(249) |
We have used Eq. (F). The other relevant matrix element is
(250) |
Time-reversal invariance relates both matrix elements. Denoting the time-reversal operator by we have
(251) | |||||
The interaction is characterized by a single parameter. For a given potential , the relevant matrix element can be calculated by expanding the axially symmetric eigenstates in terms of spherical basis functions. In our approach, is a low-energy constant and needs to be adjusted to data.
The next-to-leading-order correction of the Hamiltonian is
(254) |
Here, the “mass” matrices
(258) |
and
(261) | |||||
(264) |
enter the perturbative inversion of the mass matrix
(265) |
via
(266) |
The resulting Hamiltonian is written as in Eq. (78). Using Eq. (62) we replace the canonical momenta by angular momentum components,
(267) |
and find
(270) | |||||
(273) |
and
(279) |
With a view on Eq. (78) we note that
(280) |
and this expression is familiar to us from the leading-order Hamiltonian (65). This makes it straight forward to evaluate the next-to-leading-order corrections.
References
- Bedaque and van Kolck (2002) P. F. Bedaque and U. van Kolck, “Effective field theory for few-nucleon systems,” Annual Review of Nuclear and Particle Science 52, 339–396 (2002), nucl-th/0203055 .
- Bogner et al. (2003) S. K. Bogner, T. T. S. Kuo, and A. Schwenk, “Model-independent low momentum nucleon interaction from phase shift equivalence,” Physics Reports 386, 1 – 27 (2003).
- Epelbaum et al. (2009) E. Epelbaum, H.-W. Hammer, and Ulf-G. Meißner, “Modern theory of nuclear forces,” Rev. Mod. Phys. 81, 1773–1825 (2009).
- Bogner et al. (2010) S.K. Bogner, R.J. Furnstahl, and A. Schwenk, “From low-momentum interactions to nuclear structure,” Prog. Part. Nucl. Phys. 65, 94 – 147 (2010).
- Grießhammer et al. (2012) H. W. Grießhammer, J. A. McGovern, D. R. Phillips, and G. Feldman, “Using effective field theory to analyse low-energy compton scattering data from protons and light nuclei,” Prog. Part. Nucl. Phys. 67, 841 – 897 (2012).
- Hammer et al. (2017) H.-W. Hammer, C. Ji, and D. R. Phillips, “Effective field theory description of halo nuclei,” Journal of Physics G: Nuclear and Particle Physics 44, 103002 (2017).
- Bedaque et al. (1999) P. F. Bedaque, H.-W. Hammer, and U. van Kolck, “Renormalization of the three-body system with short-range interactions,” Phys. Rev. Lett. 82, 463–467 (1999).
- Furnstahl and Hammer (2002) R. J. Furnstahl and H.-W. Hammer, “Are occupation numbers observable?” Physics Letters B 531, 203 – 208 (2002).
- Anderson et al. (2010) E. R. Anderson, S. K. Bogner, R. J. Furnstahl, and R. J. Perry, “Operator evolution via the similarity renormalization group: The deuteron,” Phys. Rev. C 82, 054001 (2010).
- Schindler and Phillips (2009) M. R. Schindler and D. R. Phillips, “Bayesian methods for parameter estimation in effective field theories,” Ann. Phys. 324, 682 – 708 (2009).
- Furnstahl et al. (2015) R. J. Furnstahl, D. R. Phillips, and S. Wesolowski, “A recipe for eft uncertainty quantification in nuclear physics,” Journal of Physics G: Nuclear and Particle Physics 42, 034028 (2015).
- Papenbrock (2011) T. Papenbrock, “Effective theory for deformed nuclei,” Nucl. Phys. A 852, 36 – 60 (2011).
- Papenbrock and Weidenmüller (2014) T. Papenbrock and H. A. Weidenmüller, “Effective field theory for finite systems with spontaneously broken symmetry,” Phys. Rev. C 89, 014334 (2014).
- Papenbrock and WeidenmÃŒller (2015) T Papenbrock and H A WeidenmÃŒller, “Effective field theory of emergent symmetry breaking in deformed atomic nuclei,” Journal of Physics G: Nuclear and Particle Physics 42, 105103 (2015).
- Papenbrock and Weidenmüller (2016) T. Papenbrock and H. A. Weidenmüller, “Effective field theory for deformed atomic nuclei,” Physica Scripta 91, 053004 (2016).
- Chen et al. (2017) Q. B. Chen, N. Kaiser, Ulf-G. Meißner, and J. Meng, “Effective field theory for triaxially deformed nuclei,” The European Physical Journal A 53, 204 (2017).
- Chen et al. (2018) Q. B. Chen, N. Kaiser, Ulf-G. Meißner, and J. Meng, “Effective field theory for collective rotations and vibrations of triaxially deformed nuclei,” Phys. Rev. C 97, 064320 (2018).
- Chen et al. (2020) Q. B. Chen, N. Kaiser, Ulf-G. Meißner, and J. Meng, “Effective field theory for triaxially deformed odd-mass nuclei,” arXiv e-prints , arXiv:2003.04065 (2020), 2003.04065 .
- Coello Pérez and Papenbrock (2015a) E. A. Coello Pérez and T. Papenbrock, “Effective field theory for nuclear vibrations with quantified uncertainties,” Phys. Rev. C 92, 064309 (2015a).
- Coello Pérez and Papenbrock (2016) E. A. Coello Pérez and T. Papenbrock, “Effective field theory for vibrations in odd-mass nuclei,” Phys. Rev. C 94, 054316 (2016).
- Eisenberg and Greiner (1970a) Judah M. Eisenberg and Walter Greiner, Nuclear Models: Collective and Single-Particle Phenomena (North-Holland Publishing Company Ltd., London, 1970).
- Bohr and Mottelson (1975) A. Bohr and B. R. Mottelson, Nuclear Structure, Vol. II: Nuclear Deformation (W. A. Benjamin, Reading, Massachusetts, USA, 1975).
- Iachello and Arima (1987) F. Iachello and A. Arima, The Interacting Boson Model (Cambridge University Press, Cambridge, UK, 1987).
- Herzberg (1945) G. Herzberg, Molecular spectra and molecular structure, Vol. II: Infrared and Raman spectra of polyatomic molecules (Van Nostrand, New York, 1945).
- Bohr (1952) Aage Bohr, “The coupling of nuclear surface oscillations to the motion of individual nucleons,” Dan. Mat. Fys. Medd. 26, no. 14 (1952).
- Bohr and Mottelson (1953) Aage Bohr and Ben R. Mottelson, “Collective and individual-particle aspects of nuclear structure,” Dan. Mat. Fys. Medd. 27, no. 16 (1953).
- Nilsson (1955) S. G. Nilsson, “Binding states of individual nucleons in strongly deformed nuclei,” K. Dan. Vidensk. Selsk. Mat. Fys. Medd. 29, no. 16 (1955).
- Kerman (1956) A. K. Kerman, “Rotational perturbations in nuclei – application to wolfram 183,” Dan. Mat. Fys. Medd. 30, no. 15 (1956).
- Rowe (2010) David J. Rowe, Nuclear Collective Motion: Models and Theory, reprinted ed. (World Scientific, Singapore, 2010).
- Ring and Schuck (1980) P. Ring and P. Schuck, The Nuclear Many-Body Problem (Springer, Heidelberg, 1980).
- Rowe and Wood (2010) David J. Rowe and John L. Wood, Fundamentals of Nuclear Models, Vol. I: Foundational Models (World Scientific, Singapore, 2010).
- Simon (1983) Barry Simon, “Holonomy, the Quantum Adiabatic Theorem, and Berry’s Phase,” Phys. Rev. Lett. 51, 2167–2170 (1983).
- Berry (1984) Michael Victor Berry, “Quantal phase factors accompanying adiabatic changes,” Proceedings of the Royal Society of London. A. Mathematical and Physical Sciences 392, 45–57 (1984).
- Wilczek and Zee (1984) Frank Wilczek and A. Zee, “Appearance of gauge structure in simple dynamical systems,” Phys. Rev. Lett. 52, 2111–2114 (1984).
- Jackiw (1986) R. Jackiw, “Angular momentum for diatoms described by gauge fields,” Phys. Rev. Lett. 56, 2779–2780 (1986).
- Bohm et al. (1992) A. Bohm, B. Kendrick, Mark E. Loewe, and L. J. Boya, “The Berry connection and Born-Oppenheimer method,” Journal of Mathematical Physics 33, 977–989 (1992).
- Estienne et al. (2011) B. Estienne, S. M. Haaker, and K. Schoutens, “Particles in non-Abelian gauge potentials: Landau problem and insertion of non-Abelian flux,” New Journal of Physics 13, 045012 (2011).
- Nikam and Ring (1987) R. S. Nikam and P. Ring, “Manifestation of the Berry phase in diabolic pair transfer in rotating nuclei,” Phys. Rev. Lett. 58, 980–983 (1987).
- Bulgac (1990) Aurel Bulgac, “Time-dependent hartree-fock-bogoliubov approximation and nonintegrable quantum phase,” Phys. Rev. C 41, 2333–2339 (1990).
- Klein and Walet (1993) Abraham Klein and Niels R. Walet, “Quantum theory of large amplitude collective motion and the born-oppenheimer method,” Phys. Rev. C 48, 178–191 (1993).
- Nazarewicz (1994) W. Nazarewicz, “Microscopic origin of nuclear deformations,” Nuclear Physics A 574, 27 – 49 (1994).
- Hayashi (1997) Akihisa Hayashi, “Berry phase and decoupling phenomena in nuclear rotational bands,” Phys. Rev. C 56, 2326–2328 (1997).
- Chandrasekharan et al. (2008) S. Chandrasekharan, F.-J. Jiang, M. Pepe, and U.-J. Wiese, “Rotor spectra, berry phases, and monopole fields: From antiferromagnets to qcd,” Phys. Rev. D 78, 077901 (2008).
- Coello Pérez and Papenbrock (2015b) E. A. Coello Pérez and T. Papenbrock, “Effective theory for the nonrigid rotor in an electromagnetic field: Toward accurate and precise calculations of transitions in deformed nuclei,” Phys. Rev. C 92, 014323 (2015b).
- Yannouleas and Landman (2007) C. Yannouleas and U. Landman, “Symmetry breaking and quantum correlations in finite systems: studies of quantum dots and ultracold bose gases and related nuclear and chemical methods,” Rep. Prog. Phys. 70, 2067 (2007).
- Weinberg (1968) Steven Weinberg, “Nonlinear realizations of chiral symmetry,” Phys. Rev. 166, 1568–1577 (1968).
- Coleman et al. (1969) S. Coleman, J. Wess, and Bruno Zumino, “Structure of phenomenological lagrangians. i,” Phys. Rev. 177, 2239–2247 (1969).
- Callan et al. (1969) Curtis G. Callan, Sidney Coleman, J. Wess, and Bruno Zumino, “Structure of phenomenological lagrangians. ii,” Phys. Rev. 177, 2247–2250 (1969).
- Leutwyler (1994) H. Leutwyler, “Nonrelativistic effective lagrangians,” Phys. Rev. D 49, 3033–3043 (1994).
- Weinberg (1996a) S. Weinberg, The Quantum Theory of Fields, Vol. II (Cambridge University Press, Cambridge, UK, 1996).
- Brauner (2010) T. Brauner, “Spontaneous symmetry breaking and nambu-goldstone bosons in quantum many-body systems,” Symmetry 2, 609–657 (2010), arXiv:1001.5212 .
- Varshalovich et al. (1988) D. A. Varshalovich, A. N. Moskalev, and V. K. Khersonskii, Quantum theory of angular momentum (World Scientific, Singapore, 1988).
- Littlejohn and Reinsch (1997) Robert G. Littlejohn and Matthias Reinsch, “Gauge fields in the separation of rotations and internal motions in the n-body problem,” Rev. Mod. Phys. 69, 213–276 (1997).
- Román and Soto (1999) José María Román and Joan Soto, “Effective field theory approach to ferromagnets and antiferromagnets in crystalline solids,” International Journal of Modern Physics B 13, 755–789 (1999).
- Hofmann (1999) Christoph P. Hofmann, “Spin-wave scattering in the effective lagrangian perspective,” Phys. Rev. B 60, 388–405 (1999).
- Bär et al. (2004) O. Bär, M. Imboden, and U.-J. Wiese, “Pions versus magnons: from qcd to antiferromagnets and quantum hall ferromagnets,” Nuclear Physics B 686, 347 – 376 (2004).
- Kämpfer et al. (2005) F. Kämpfer, M. Moser, and U.-J. Wiese, “Systematic low-energy effective theory for magnons and charge carriers in an antiferromagnet,” Nuclear Physics B 729, 317 – 360 (2005).
- Lepage (1997) G. P. Lepage, “How to renormalize the Schrodinger equation,” in Nuclear physics. Proceedings, 8th Jorge Andre Swieca Summer School, Sao Jose dos Campos, Campos do Jordao, Brazil, January 26-February 7, 1997 (1997) pp. 135–180, nucl-th/9706029 .
- Scaldeferri et al. (1997) K. A. Scaldeferri, D. R. Phillips, C.-W. Kao, and T. D. Cohen, “Short-range interactions in an effective field theory approach for nucleon-nucleon scattering,” Phys. Rev. C 56, 679–688 (1997).
- Beane et al. (1998) S.R. Beane, T.D. Cohen, and D.R. Phillips, “The potential of effective field theory in nn scattering,” Nuclear Physics A 632, 445 – 469 (1998).
- Kirscher et al. (2010) J. Kirscher, H. W. Grießhammer, D. Shukla, and H. M. Hofmann, “Universal correlations in pion-less eft with the resonating group method: Three and four nucleons,” The European Physical Journal A 44, 239–256 (2010).
- Lensky et al. (2016) Vadim Lensky, Michael C. Birse, and Niels R. Walet, “Description of light nuclei in pionless effective field theory using the stochastic variational method,” Phys. Rev. C 94, 034003 (2016).
- Capel et al. (2018) P. Capel, D. R. Phillips, and H.-W. Hammer, “Dissecting reaction calculations using halo effective field theory and ab initio input,” Phys. Rev. C 98, 034610 (2018).
- Eisenberg and Greiner (1970b) Judah M. Eisenberg and Walter Greiner, Nuclear Models: Collective and Single-Particle Phenomena (North-Holland Publishing Company Ltd., London, 1970).
- Kuratsuji and Iida (1985) Hiroshi Kuratsuji and Shinji Iida, “Effective Action for Adiabatic Process: Dynamical Meaning of Berry and Simon’s Phase,” Progress of Theoretical Physics 74, 439–445 (1985).
- Mead and Truhlar (1979) C. Alden Mead and Donald G. Truhlar, “On the determination of Born-Oppenheimer nuclear motion wave functions including complications due to conical intersections and identical nuclei,” The Journal of Chemical Physics 70, 2284–2296 (1979).
- Rho (1994) Mannque Rho, “Cheshire cat hadrons,” Physics Reports 240, 1 – 142 (1994).
- Fierz (1944) M. Fierz, “Zur Theorie magnetisch geladener Teilchen,” Helvetica Physica Acta 17, 27 (1944).
- Wu and Yang (1976) Tai Tsun Wu and Chen Ning Yang, “Dirac monopole without strings: Monopole harmonics,” Nuclear Physics B 107, 365 – 380 (1976).
- Zhang and Papenbrock (2013) Jialin Zhang and T. Papenbrock, “Rotational constants of multi-phonon bands in an effective theory for deformed nuclei,” Phys. Rev. C 87, 034323 (2013).
- Malmskog et al. (1971) S. G. Malmskog, V. Berg, B. Fogelberg, and A. Bäcklin, “On the low-energy band structure in 187os,” Nuclear Physics A 166, 573 – 593 (1971).
- Morgen et al. (1973) P. Morgen, B. S. Nielsen, J. Onsgaard, and C. Søndergaard, “The level structure of 187os,” Nuclear Physics A 204, 81 – 96 (1973).
- Sodan et al. (1975) H. Sodan, W. D. Fromm, L. Funke, K. H. Kaun, P. Kemnitz, E. Will, G. Winter, and J. Berzins, “Rotational bands in 185os and 187os,” Nuclear Physics A 237, 333 – 353 (1975).
- Noether (1918) Emmy Noether, “Invariante Variationsprobleme,” Nachrichten von der Königlichen Gesellschaft der Wissenschaften zu Göttingen, Mathematisch-Physikalische Klasse 1918, 235–257 (1918).
- Nauts and Gatti (2010) André Nauts and Fabien Gatti, “Unusual commutation relations in physics,” American Journal of Physics 78, 1365–1375 (2010).
- Leutwyler (1997) H. Leutwyler, “Phonons as goldstone bosons,” Helv. Phys. Acta 70, 275–286 (1997), arXiv:hep-ph/9609466 [hep-ph] .
- Weinberg (1996b) S. Weinberg, The Quantum Theory of Fields, Vol. II (Cambridge University Press, Cambridge, UK, 1996).
- Gasser and Leutwyler (1988) J. Gasser and H. Leutwyler, “Spontaneously broken symmetries: Effective lagrangians at finite volume,” Nuclear Physics B 307, 763 – 778 (1988).
- Wilczek and Shapere (1989) F. Wilczek and A. Shapere, Geometric Phases in Physics (World Scientific, Singapore, 1989) https://www.worldscientific.com/doi/pdf/10.1142/0613 .
- Lai et al. (1998) Shan-Tao Lai, Pancracio Palting, Ying-Nan Chiu, and Harris J. Silverstone, “On the summations involving wigner rotation matrix elements,” Journal of Mathematical Chemistry 24, 123 – 132 (1998).
- Dray (1986) Tevian Dray, “A unified treatment of wigner d functions, spin-weighted spherical harmonics, and monopole harmonics,” Journal of Mathematical Physics 27, 781–792 (1986).