Effective Lagrangian for the Macroscopic Motion of Weyl Fermions in 3He-A
Abstract
1. Introduction
2. The Emergence and Characteristics of the 3He-A Superfluid
2.1. The Action of 3He with Omission of Spin–Orbit Interaction
2.2. Vacuum of 3He-A in the London Limit for Inhomogeneous Fields
3. Low-Energy Effective Theory with Emergent Relativistic Invariance
3.1. Fermionic Action Near the Fermi Points
- (1)
- The vierbein and its inverse are matrix valued due to the term . We define scalar-valued vierbeins byWe use bold letters in the case of matrix-valued vierbein comprising the term and the unbolded notation, if it is absent. Moreover, the scalar vierbein determinants coincide with that of the matrix-valued vierbein . The motivation to consider these scalar vierbeins arises due to the consideration of the eigenspaces of the matrix-valued operator . We subsequently suppress the superscript on the scalar-valued vierbein and consider it only implicitly. Further below, we will consider the choice . We define the projection operatorsWe have the common notation . The normalized eigenspinors of this operator for areWe introduce the two-component spinors as follows:(In the following, we will omit the symbol ⊗ of the tensor product for brevity.) ThenWe may further employ the projection operators () in order to represent the 8-component Dirac spinors in terms of the 4-component spinors
- (2)
- The vierbein (as well as ) is not orthonormal with respect to the Minkowski metric but instead fulfillsThis metric is a natural measure of distance within superfluid 3He-A. We use and its inverse to raise or lower spacetime indices (Greek letters) and and its inverse to raise or lower Lorentz indices (Latin letters a, b, c …). Spatial spacetime indices are labeled by Latin letters i, j, k …. The spacetime we are working on is flat as a consequence of the constancy of . We will furthermore work with the definition for the -symbol, where the indices refer to the local Lorentz frame.
- (3)
- The action has vanishing spin connection gauge field . Its presence is required in the standard relativistic theory in order to ensure local Lorentz invariance. Instead we are only given the gauge field of translations with its curvatureThe minus sign is in line with the standard definitionUsing we obtain and . This leads toThe tensor field is also known as the torsion tensor field. In absence of the spin connection, we will not require that the vierbein is covariantly constant in general. In this context, the covariant constancy of the vierbein is equivalent to a constant vierbein. The action then features global translation and Lorentz invariance. We will nevertheless introduce the spin connection in order to derive the spin tensor below. In the relativistic theory of Dirac fermions, the spin connection enters via the covariant derivative (which is diagonal in the internal spin space due to )(note that our overline does not comprise ). In the relativistic theory of Weyl fermions, the spin connection enters via the covariant derivativeThe above expressions hold in the same way when written in terms of and . We will modify the definition in the latter case, though by replacing simultaneously by . This definition will be motivated shortly and gives rise to a different spin tensor. This will be remarked again further below.
3.2. A Reparametrization—Universal Vierbein Field and Spin Connection Gauge Field
4. The Zubarev Statistical Operator Formalism
4.1. Essentials of the Zubarev Statistical Operator Method
4.2. The Zubarev Statistical Operator Within Superfluid 3He-A
4.2.1. Equations of Motion
4.2.2. Conserved and Non-Conserved Currents
4.2.3. Stationarity of the Statistical Operator: Assumption of Global Thermodynamic Equilibrium
5. Effective Lagrangian in the Presence of Macroscopic Motion Within the Path Integral Formulation
- 1.
- 2.
- 3.
- We define Grassmann-valued fields as related to by the same expression that relates and a. One can check easily that (or equivalently for the left-handed fields).
6. Thermodynamic Equilibrium Solutions for the Vierbein
- (1)
- . This case is equivalent to a vanishing Lie derivative of the vierbein along the Killing vector field .
- (2)
- (supplemented by Equation (160) in the case of inhomogeneous d or non-vanishing ). This case implies homogeneity of the vierbein in space and time and moreover that .
- (1)
- Pure (integer) mass vortices: We consider a stationary setup with vanishing translational velocity , acceleration but nonzero rigid rotation around the z-axis with with angular velocity and inverse temperature . It is convenient to introduce cylindrical coordinates expressed through Cartesian coordinates byWe denote Cartesian unit vectors by , and and those of the cylindrical reference frame by , and , respectively. Then the pure mass vortices are given byWe also have . The pure mass vortex is a global thermodynamic equilibrium solution in the presence of both - and -components, as they are compatible with Equations (146) and (160). Consider now a slight misalignment of the rotation axis of macroscopic motion from that of the vortex axis . We have , which is compatible with Equation (160) since and also with Equation (148). This means that the global thermodynamic equilibrium is achieved for any relative orientation of vortex and macroscopic motion rotation axis. This is a consequence of the fact that not the Lie derivative of the torsion tensor field along the frigidity vector field is supposed to vanish in global thermodynamic equilibrium but only its antisymmetrization. If d becomes dipole unlocked with l such that , the global thermodynamic equilibrium is achievable if and only if the vortex axis is aligned with the rotation axis of macroscopic motion.We consider now the Lagrangian and the equations of motion in the case of alignment of the vortex axis and the rotation axis of macroscopic motion with dipole locked d. It suffices to treat the -component since both components are not coupled for homogeneous d. We further restrict to the left-handed fermions. The case of right-handed fermions may be obtained by setting , implying a helicity flip. The Lagrangian of Equation (197) for our case is given by (with )The pure mass vortices we are considering have an infinitely thin core. We are interested in the normalizable bound states. The four terms in Equation (216) cannot vanish independently, unless either or . We therefore require (unless either or , but this is not compatible with normalizability).In the case where both and , the eigenvalue problem with becomesDefine such that for even and for odd . Equation (217) then becomesNotice that we reduce the problem to the case without vortex by this index shift. We will proceed by assuming a solution of the formThis leads to the new decoupled differential equationsWe require and define . This requirement is necessary in order to interpret the eigenfunctions as bound states. Together with the coordinate redefinition solutions to these equations are Bessel functions of the first kind with as well as and . Equation (218) then becomes fully algebraic. Employing the relationsThe eigenvalues and amplitudes are finally given bySince only in general, we introduce with such that . We obtain the solutions of the right-handed fermions by the replacement which changes the eigenstates, while the energy levels remain invariant. We thus have a degeneracy of the four distinct degrees of freedom spanned by the helicity (or chirality) eigenstates with eigenvalues and the eigenstates with intrinsic spin eigenvalues . The original number of degrees of freedom are restricted from eight to four due to the constraint Equation (25) (see also Equations (A5) and (A6) in Appendix A).We proceed to calculate the grand canonical potential for our system of fermions at finite temperature T, chemical potential , and angular velocity . The chemical potential is due to the axial charge since the vector charge of the Lagrangian in Equation (203) vanishes after restriction to the constrained variables. The system is necessarily bound to be finite by the causality constraint with R as the transverse radius. We will then need to introduce boundary conditions for the spinors. We choose MIT bag boundary conditionsThese boundary conditions imply a mixing of the chiral components according toThe final equality within Equation (227) implies the quantization of . We will call its quantized values . Since the MIT bag boundary conditions imply a relation between the chiral components, we will restore a factor of two in the number of degrees of freedom counting. The (vacuum subtracted and therefore renormalized) pressure (or the negative grand canonical potential) readsIn ordinary relativistic units, we haveIn the following we use the units of Table 1.We plot the pressure p (in the units of ) as well as the particle number densityA numerical evaluation of the pressure, energy density and entropy density over temperature ranges of five and six e-folds yields the results presented in Figure 3, Figure 4 and Figure 5 pressure, energy, entropy, respectively.In the upper plots we compare the calculated values with the expected asymptotic high-temperature limitsIn the lower plots, we compare the topologically distinct cases and by considering the relative difference of the two cases for pressure, energy density and entropy density, respectively. While at high temperatures the presence of a vortex has only a minor effect on the thermodynamic quantities, the low temperature limit exposes the effect of a vortex quite visibly since at low temperatures only a small number of discrete modes is thermodynamically accessible. An increase in temperature enhances the number of discrete modes participating in thermodynamic fluctuations. The pressure, energy density, and entropy density become significantly reduced in the presence of a vortex. The relative suppression decreases with chemical potential and angular velocity. If both the chemical potential and angular velocity are large, the presence of a vortex is basically irrelevant to the thermodynamics. For low angular velocities and intermediate-to-large chemical potentials, the entropy density even exhibits an enhancement in the presence of a vortex at low temperatures, which is observed neither for the pressure nor the entropy densities, however.A numerical evaluation of the entropy per particle as well as the angular momentum per particle over a temperature range of six e-folds is presented in the upper plots of Figure 6 and Figure 7, respectively. We furthermore show, in line with the previous plots of the pressure, energy density, and entropy density, the relative difference of particle number densities and angular momentum densities for the two topologically distinct cases and over a range of five e-folds in temperature. The latter (lower) plots exhibit the same patterns as outlined before in the cases of the pressure and energy density, so we proceed to discuss the upper plots.At high temperatures, the entropy per particle converges to a finite value which is larger for lower angular velocity but not dependent on the chemical potential. Towards lower temperatures, the entropy per particle develops a strong dependence on the chemical potential. For very small particle density (), it increases for decreasing temperature, while it increases for very large particle number density (). It is monotonously decreasing with the chemical potential. In contrast, the entropy per particle depends only weakly on the angular velocity without any specific properties regarding monotony. At high temperatures the angular momentum per particle converges to a finite value as well. The asymptotic region implied by the convergence is reached at smaller temperatures for large angular velocity. For fixed temperatures, the angular momentum per particle increases both for increasing angular velocity and chemical potential. Towards low temperatures, a strong dependence of the angular momentum per particle both on the chemical potential and angular velocity arises. It decreases monotonously with temperature.The consideration of the number of degrees of freedom mentioned further above is mostly irrelevant for the discussed plots, as the considered ratios of thermodynamic quantities are independent of the number of degrees of freedom.
- (2)
- Disclinations: We again assume stationarity with vanishing translational velocity and acceleration . We would like to consider once more the situation with nonzero rigid rotation of macroscopic motion. A radial (tangential) disclination in 3He-A is defined byThe vector d is again dipole locked with l. We find (, ) andNow is not naturally vanishing for disclinations which implies for the Lorentz indexed vector fields
- (3)
- Fractional vortices: We consider fractional (or spin mass) vortices corresponding to the situation outlined above in the discussion of pure mass vortices but now with such that . In order for the order parameter of the 3He-A-phase to be single-valued, d can no longer be dipole locked with l but instead fulfillsTherefore, (or equivalently ) is naturally nonzero
7. Discussion
- 1.
- Motion with constant four-velocity . Correspondingly, in this case, the temperature and chemical potential are constant as functions of time.
- 2.
- Rigid rotation with constant angular velocity . In this case, temperature becomes a function of spatial coordinates. The whole theory becomes ill-defined at distances larger than from the rotation axis. This means that we can use the Zubarev statistical operator for the case when , where R is the size (transverse extension) of the considered system. This admits, in particular, the possibility of rotation with relativistic velocities. The inequality is valid by causality. Any physical system will disintegrate (or be ripped apart) before reaching the limit implied by causality.If rotation is along the z-axis, then we have
- 3.
- Accelerated motion with constant acceleration vector a. The acceleration a appears as the thermodynamically conjugated quantity to the boost operator. The interpretation of the theory in terms of the four-velocity becomes ill-defined at times . This situation is avoided in a physical system, as maintaining the finite acceleration a for times larger than implies an injection of an infinite amount of energy. However, the spatial size of the corresponding system is not limited.In the case when acceleration is along the x-axis, we have
- 4.
- The combination of the three types of motion explained above is also admitted for the thermal equilibrium.
- 1.
- Uniform linear motion along the x-direction with constant four velocityIn this case is constant as well
- 2.
- Rigid rotation with constant angular velocity around the z-axisThen we have
- 3.
- The initially accelerated motion with acceleration a along the x-axisFor the choice we haveOne can see that in this case (especially for accelerated motion) the effective Lagrangian is especially simple. For accelerated motion it is reduced to the Lagrangian of the system remaining at rest. The only effect of acceleration is manifested through the temperature depending on the spatial coordinates.
8. Conclusions
Author Contributions
Funding
Data Availability Statement
Acknowledgments
Conflicts of Interest
Appendix A. Equations of Motion and Canonical Formalism for Nambu–Gorkov Spinors Without Doubling the Degrees of Freedom
Appendix A.1. Equations of Motion
Appendix A.2. Canonical Quantization
Appendix B. Canonical Quantization and Fermion Doubling
Appendix C. Energy–Momentum Tensor and Spin Tensor Non-Conservation
Appendix D. System of Units
energy | temperature | ||
momentum | pressure | ||
mass | entropy density | ||
time | particle number density | ||
position | angular momentum density |
energy | temperature | ||
momentum | pressure | ||
mass | entropy density | ||
time | particle number density | ||
position | angular momentum density |
References
- Volovik, G.E. The Universe in a Helium Droplet; Clarendon Press: Oxford, UK, 2003. [Google Scholar]
- Vollhardt, D.; Wölfle, P. The Superfluid Phases of Helium 3; Taylor and Francis: London, UK, 1990. [Google Scholar]
- Volovik, G.; Zubkov, M. Emergent weyl spinors in multi-fermion systems. Nucl. Phys. B 2014, 881, 514–538. [Google Scholar] [CrossRef]
- Volovik, G.E.; Vachaspati, T. Aspects of 3He and the standard electroweak model. Int. J. Mod. Phys. B 1996, 10, 471–521. [Google Scholar] [CrossRef]
- Volovik, G.E.; Khazan, M.V. Dynamics of the A-phase of 3He at low pressure. JETP 1983, 55, 551. [Google Scholar]
- Volovik, G.E. Symmetry in Superfluid 3He. Mod. Probl. Condens. Matter Sci. 1990, 26, 27–134. [Google Scholar]
- Nambu, Y. Fermion-boson relations in BCS type theories. Phys. D Nonlinear Phenom. 1985, 15, 147–151. [Google Scholar] [CrossRef]
- Volovik, G.E.; Zubkov, M.A. Scalar excitation with leggett frequency in 3He-B and the 125 gev higgs particle in top quark condensation models as pseudo-goldstone bosons. Phys. Rev. D 2015, 92, 055004. [Google Scholar] [CrossRef]
- Volovik, G.E.; Zubkov, M.A. Higgs bosons in particle physics and in condensed matter. J. Low Temp. Phys. 2014, 175, 486–497. [Google Scholar] [CrossRef]
- Volovik, G.E.; Zubkov, M.A. Nambu sum rule and the relation between the masses of composite higgs bosons. Phys. Rev. D 2013, 87, 075016. [Google Scholar] [CrossRef]
- Zubkov, M.A. Wigner transformation, momentum space topology, and anomalous transport. Ann. Phys. 2016, 373, 298–324. [Google Scholar] [CrossRef]
- Zhang, C.X.; Zubkov, M.A. Influence of interactions on the anomalous quantum hall effect. J. Phys. A Math. Theor. 2021, 53, 195002. [Google Scholar] [CrossRef]
- Zhang, C.X.; Zubkov, M.A. Influence of interactions on integer quantum hall effect. Ann. Phys. 2022, 444, 169016. [Google Scholar] [CrossRef]
- Zubkov, M.A.; Wu, X. Topological invariant in terms of the green functions for the quantum hall effect in the presence of varying magnetic field. Ann. Phys. 2020, 418, 168179. [Google Scholar] [CrossRef]
- Suleymanov, M.; Zubkov, M.A.; Zhang, C.X. Hall conductivity as the topological invariant in magnetic Brillouin zone, preprint. arXiv 2021, arXiv:2112.03974. [Google Scholar]
- Nieh, H.; Yan, M. Quantized dirac field in curved riemann-cartan background. I. symmetry properties, green’s function. Ann. Phys. 1982, 138, 237–259. [Google Scholar] [CrossRef]
- Huang, Z.-M.; Han, B.; Stone, M. Nieh-yan anomaly: Torsional landau levels, central charge, and anomalous thermal hall effect. Phys. Rev. B 2020, 101, 125201. [Google Scholar] [CrossRef]
- Khaidukov, Z.V.; Zubkov, M. Chiral torsional effect. JETP Lett. 2018, 108, 670–674. [Google Scholar] [CrossRef]
- Nissinen, J.; Volovik, G.E. On thermal nieh-yan anomaly in topological weyl materials. JETP Lett. 2019, 110, 789–792. [Google Scholar] [CrossRef]
- Nissinen, J.; Volovik, G. Thermal nieh-yan anomaly in weyl superfluids. Phys. Rev. Res. 2020, 2, 033269. [Google Scholar] [CrossRef]
- Nissinen, J. Emergent spacetime and gravitational nieh-yan anomaly in chiral p+ ip weyl superfluids and superconductors. Phys. Rev. Lett. 2020, 124, 117002. [Google Scholar] [CrossRef]
- Laurila, S.; Nissinen, J. Torsional landau levels and geometric anomalies in condensed matter weyl systems. Phys. Rev. B 2020, 102, 235163. [Google Scholar] [CrossRef]
- DZubarev, N.; Prozorkevich, A.V.; Smolyanskii, S.A. Derivation of nonlinear generalized equations of quantum relativistic hydrodynamics. Theor. Math. Phys. 1979, 40, 821–831. [Google Scholar] [CrossRef]
- Selch, M.; Abramchuk, R.A.; Zubkov, M. Effective lagrangian for the macroscopic motion of fermionic matter. Phys. Rev. D 2024, 109, 016003. [Google Scholar] [CrossRef]
- Alonso, V.; Popov, V.N. Functional for the hydrodynamic action and the bose spectrum of superfluid fermi systems of the He3 type. Zh. Eksp. Teor. Fiz. 1977, 73, 1445. [Google Scholar]
- Brusov, P.N.; Popov, V.N. Stability of the bose spectrum of superfluid systems of the He3 type. Zh. Eksp. Teor. Fiz. 1980, 78, 234–245. [Google Scholar]
- Brusov, P.N.; Popov, V.N. Nonphonon branches of the bose spectrum in the B phase of systems of the He3 type. JETP 1980, 51, 1217–1222. [Google Scholar]
- Brusov, P.N.; Popov, V.N. Zero-phonon branches of the bose spectrum in the A phase of a system of the He3 type. JETP 1980, 52, 945–949. [Google Scholar]
- Brusov, P.; Brusov, P. Collective Excitations in Unconventional Superconductors and Superfluids; World Scientific Publishing Co. Pte. Ltd.: Singapore, 2010. [Google Scholar]
- Buzzegoli, M.; Grossi, E.; Becattini, F. General equilibrium second-order hydrodynamic coefficients for free quantum fields. J. High Energy Phys. 2017, 2017, 91. [Google Scholar] [CrossRef]
- Becattini, F.; Buzzegoli, M.; Palermo, A. Exact equilibrium distributions in statistical quantum field theory with rotation and acceleration: Scalar field. J. High Energy Phys. 2021, 2, 101. [Google Scholar] [CrossRef]
- Buzzegoli, M. Thermodynamic equilibrium of massless fermions with vorticity, chirality and electromagnetic field. Lect. Notes Phys. 2021, 987, 59–93. [Google Scholar]
- Bravina, L.V.; Bugaev, K.A.; Vitiuk, O.; Zabrodin, E.E. Transport Model Approach to and - Polarization in Heavy-Ion Collisions. Symmetry 2021, 13, 1852. [Google Scholar] [CrossRef]
- Selch, M.; Zubkov, M.A.; Pramanik, S.; Lewkowicz, M. Non-renormalization of the fractional quantum hall conductivity by interactions. arXiv 2025, arXiv:2502.04047. [Google Scholar]
- Yamamoto, A.; Hirono, Y. Lattice QCD in rotating frames. Phys. Rev. Lett. 2013, 111, 081601. [Google Scholar] [CrossRef] [PubMed]
- Braguta, V.V.; Kotov, A.Y.; Kuznedelev, D.D.; Roenko, A.A. Study of the Confinement/Deconfinement Phase Transition in Rotating Lattice SU(3) Gluodynamics. Pisma Zh. Eksp. Teor. Fiz. 2020, 112, 6–12. [Google Scholar] [CrossRef]
- Braguta, V.V.; Kotov, A.Y.; Kuznedelev, D.D.; Roenko, A.A. Influence of relativistic rotation on the confinementdeconfinement transition in gluodynamics. Phys. Rev. D 2021, 103, 094515. [Google Scholar] [CrossRef]
- Chernodub, M.N.; Goy, V.A.; Molochkov, A.V. Inhomogeneity of a rotating gluon plasma and the Tolman-Ehrenfest law in imaginary time: Lattice results for fast imaginary rotation. Phys. Rev. D 2023, 107, 114502. [Google Scholar] [CrossRef]
- Braguta, V.V.; Kudrov, I.E.; Roenko, A.A.; Sychev, D.A.; Chernodub, M.N. Lattice Study of the Equation of State of a Rotating Gluon Plasma. JETP Lett. 2023, 117, 639–644. [Google Scholar] [CrossRef]
- Braguta, V.V.; Chernodub, M.N.; Roenko, A.A.; Sychev, D.A. Negative moment of inertia and rotational instability of gluon plasma. arXiv 2023, arXiv:2303.03147. [Google Scholar] [CrossRef]
- Morales-Tejera, S.; Landsteiner, K. Out of equilibrium chiral vortical effect in holography. Phys. Rev. D 2020, 102, 106020. [Google Scholar] [CrossRef]
- Prokhorov, G.Y.; Teryaev, O.V.; Zakharov, V.I. Novel phase transition at the Unruh temperature. arXiv 2023, arXiv:2304.13151. [Google Scholar]
- Khakimov, R.V.; Prokhorov, G.Y.; Teryaev, O.V.; Zakharov, V.I. Unruh effect in curved space-time and hydrodynamics. arXiv 2023, arXiv:2308.08647. [Google Scholar]
- Chernodub, M.; Pochinok, A.S.; Molochkov, A.; Stepanov, D.V.; Goy, V. Gluon matter under weak acceleration: Lattice results. In Proceedings of the 8th International Conference on Chirality, Vorticity, and Magnetic Field in Quantum Matter, West University of Timișoara, Timisoara, Romania, 22–26 July 2024. [Google Scholar]
- Hayata, T.; Hidaka, Y.; Noumi, T.; Hongo, M. Relativistic hydrodynamics from quantum field theory on the basis of the generalized gibbs ensemble method. Phys. Rev. D 2015, 92, 065008. [Google Scholar] [CrossRef]
- Hongo, M. Path-integral formula for local thermal equilibrium. Ann. Phys. 2017, 383, 1–32. [Google Scholar] [CrossRef]
- Volovik, G.E. Exotic Properties of Superfluid Helium 3; World Scientific: Singapore, 1992. [Google Scholar]
- Fujii, K.; Nishida, Y. Low-energy effective field theory of superfluid 3He-B and its gyromagnetic and hall responses. Ann. Phys. 2018, 395, 170–182. [Google Scholar] [CrossRef]
- Furusawa, T.; Fujii, K.; Nishida, Y. Hall viscosity in the A phase of superfluid 3He. Phys. Rev. B 2021, 103, 064506. [Google Scholar] [CrossRef]
energy | temperature | ||
momentum | pressure | ||
mass | entropy density | ||
time | particle number density | ||
position | angular momentum density |
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Selch, M.; Zubkov, M. Effective Lagrangian for the Macroscopic Motion of Weyl Fermions in 3He-A. Symmetry 2025, 17, 1045. https://doi.org/10.3390/sym17071045
Selch M, Zubkov M. Effective Lagrangian for the Macroscopic Motion of Weyl Fermions in 3He-A. Symmetry. 2025; 17(7):1045. https://doi.org/10.3390/sym17071045
Chicago/Turabian StyleSelch, Maik, and Mikhail Zubkov. 2025. "Effective Lagrangian for the Macroscopic Motion of Weyl Fermions in 3He-A" Symmetry 17, no. 7: 1045. https://doi.org/10.3390/sym17071045
APA StyleSelch, M., & Zubkov, M. (2025). Effective Lagrangian for the Macroscopic Motion of Weyl Fermions in 3He-A. Symmetry, 17(7), 1045. https://doi.org/10.3390/sym17071045