Microscopic theory of pair density waves in spin-orbit coupled Kondo lattice
Abstract
We demonstrate that the discommensuration between the Fermi surfaces of a conduction sea and an underlying spin liquid provides a natural mechanism for the spontaneous formation of pair density waves. Using a recent formulation of the Kondo lattice model which incorporates a Yao Lee spin liquid proposed by the authors, we demonstrate that doping away from half-filling induces finite-momentum electron-Majorana pair condensation, resulting in amplitude-modulated PDWs. Our approach provides a precise, analytically tractable pathway for understanding the spontaneous formation of PDWs in higher dimensions and offers a natural mechanism for PDW formation in the absence of a Zeeman field.
pacs:
PACS TODOIntroduction. Spatially modulated pair condensates, or pair density wavesAgterberg et al. (2020) (PDWs) were originally envisioned by Fulde, FerrellFulde and Ferrell (1964), Larkin and OvchinikovLarkin and Ovchinnikov (1964)(FFLO), as a high-field superconducting response to the Zeeman splitting of the Fermi surface. The recent observation of PDWs via scanning tunneling microscopyRanderia et al. (2016) measurements in cuprateHücker et al. (2011), iron-based Zhao et al. (2023); Liu et al. (2023) and heavy fermion superconductorsGu et al. (2023) has sparked a resurgence of interest in this phenomenon. Crucially, in each of these unconventional superconductors the PDW develops without a magnetic field. PDWs naturally form as a response of a uniform pair condensate to a charge or spin density wave(CDW)Agterberg et al. (2020); Fradkin et al. (2015). However, the recent PDW observation in the novel triplet superconductor, UTe2Gu et al. (2023) found that the PDW survives the melting of CDW orderAishwarya et al. (2024), suggesting that in this case, the PDW is a primary order parameter, developing without the aid of a charge or spin density wave. These observations motivate us to seek a natural mechanism for the spontaneous development of pair density wave order.
The condensation of a PDW requires a pair susceptibility that peaks at a finite momentum. This does not occur naturally in a BCS superconductorBardeen et al. (1957), where the logarithmic divergence of the uniform pair susceptibility is cut off by the momentum, or more precisely by , where is the Fermi velocity and the wavevector of the PDW. In a BCS superconductors, a Zeeman field polarizes the Fermi surfaces, cutting off the zero-momentum instability and favoring a finite-momentum peak in the pair susceptibility, leading to an FFLO-like pair density wavesFulde and Ferrell (1964); Larkin and Ovchinnikov (1964). While experiment suggests that pair density waves can spontaneously develop in unconventional superconductors even in the absence of a magnetic field, there is currently no current consensus on the underlying mechanism.
In this paper we explore PDW formation in the context of the Kondo lattice model for heavy fermions. The current authors (CPT) have recently proposed a three dimensional, solvable model in which the Kondo screening of a spin liquid induces superconductivity in the surrounding conduction seaColeman et al. (2022); Tsvelik and Coleman (2022). The underlying spin liquid is a three dimensional generalization of the spin-orbital derivative of the well-known Kitaev spin liquid, known as the Yao-Lee spin liquidYao and Lee (2011)). At half-filling and in the weak coupling limit, Kondo coupling between the spin liquid and the conduction electrons induces odd-frequency triplet superconductivity.
In this letter, we demonstrate that our modelColeman et al. (2022) provides a natural mechanism for the emergence of incommensurate pair-density waves when the system is doped away from half-filling. In particular, its weak coupling description offers a long-sought pathway to understanding the spontaneous formation of pair-density waves beyond one dimension Agterberg et al. (2020) in the absence of a field.

At half-filling, the model exhibits nesting between electron, hole, and Majorana Fermi surfaces, leading to a logarithmic instability into an electron-Majorana pair condensate under infinitesimal Kondo coupling. Our key result is that doping the system away from half-filling, through a shift in the chemical potential, modifies the electron and hole Fermi surfaces while leaving the Majorana Fermi surface unchanged. This imbalance induces a finite-momentum FFLO-like electron-Majorana condensation, giving rise to amplitude-modulated pair-density waves.
A key advantage of our model over general parton theories is that it does not rely on an approximate Gutzwiller projection to enforce the single-occupancy constraint by mean-field methods. Instead, in the Yao-Lee spin liquid, this constraint is inherently encoded in the Majorana representation, offering a more precise understanding of how the chemical potential influences the order parameter.
Model. The CPT modelColeman et al. (2022) is defined on the hyperoctagon lattice (Fig. 1), Kondo-couples conduction electrons to the spins of a Yao-Lee spin liquid. Each site possesses three degrees of freedom: conduction electrons, localized spins, and localized orbitals, leading to a Hamiltonian with three components, where describes the nearest-neighbor hopping of the conduction electrons, captures the Yao-Lee spin-spin interaction, and couples the conduction sea to the Yao-Lee spin liquid through the Kondo interaction,
(1) | ||||
Yao-Lee spin liquid: The Yao-Lee term involves an anisotropic Kitaev interaction between the orbitals ’s components on nearest-neighbor sites , decorated by a Heisenberg interaction between the spins . The Hamiltonian can be significantly simplified through fermionization, where the commutation relations of the orbital and spin operators are encoded by expressing the orbitals as and the spins as , in terms of Majorana fermions and . The physical Hilbert space is constrained by , which commutes with the Hamiltonian.
The orbital frustration in the Yao-Lee Hamiltonian induces a Majorana fractionalization of spins, which couple to static gauge fields , , An Ising deconfinement transition occurs at a finite temperature , where the gauge fields freeze into a flux-free configuration. This leads to a four-band Majorana Hamiltonian on the hyperoctagon lattice. At low energies, the Yao-Lee spin liquid can be effectively described by a single-band with dispersion , where
(2) |
where, and . The projected Hamiltonian is then expressed as,
(3) |
where the momentum sum takes place over the cubic Majorana Brillouin zone \mancube.

Conduction electrons: Similarly, the low energy physics of the conduction band following a non-singular gauge transformationColeman et al. (2022), can be described by a projected single band , with taking the form (2). Further, the conduction sea can be restricted to \mancube, by rewriting the conduction Hamiltonian in terms of the Balian-Werthammer spinor as,
(4) |
Kondo interaction: Below the Ising phase boundary of Yao-Lee spin liquid, the Majorana deconfinement of spins enables the re-expression of Kondo interactions in terms of Majorana fermions and conduction electrons. This reformulation facilitates an analytic weak coupling mean-field treatment of the Kondo interaction without relying on Gutzwiller projection. In this representation, the spins are expressed as , allowing the Kondo interaction to be rewritten as,
(5) |
can be decoupled via a Hubbard-Stratonovich transformation in terms of a spinor order parameter field , which is solved self-consistently. Expressing the conduction electrons and the spinor order parameter using the Balian-Werthamer representation, and , respectively, provides a compact mean-field solution for the Kondo interaction.
(6) |
In previous work Coleman et al. (2022), we showed that at half-filling, the model exhibits a logarithmic singularity in the electron-Majorana pair susceptibility, leading to odd-frequency triplet superconductivity.
Pair density waves. We now show that doping this analytically tractable Kondo lattice model provides a natural mechanism for the spontaneous formation of a pair density wave (PDW). Here, the chemical potential plays a role analogous to the Zeeman field in conventional superconductors. Specifically, the addition of a chemical potential splits the electron and hole Fermi surfaces while leaving the Majorana Fermi surface unaffected (Fig 2a). This then causes the fractionalized spinor order parameter to acquire finite momentum (Fig. 2b), resulting in the formation of a PDW state. In this scenario, the Fourier transform of the Kondo interaction with finite momentum order parameter is given by,
(7) |
The tendency to form a finite-momentum spinor order arises from the static electron-Majorana pairing susceptibility (Fig. 3) being maximized at a finite momentum ,
(8) |
Which is expressed in terms of the conduction and Majorana Green’s functions, , and respectively, with representing the orientation of the spinor order. Further simplifications to trace conditions are made by noting that forms a projection operator, which decouples a conduction Majorana as a result of Kondo screening.
Following simplifications, outlined in the supplementary material, we can rewrite the susceptibility for the hyperoctagon lattice in the following closed form,
(9) |

Here, is the density of states at the Fermi surface, and is the Fermi velocity of the conduction sea, which, for the hyperoctagon lattice, depends on the orientation . The logarithmic instability of the electron-Majorana susceptibility in equation (9) is cut off by the chemical potential, which splits the nesting between the conduction sea and the Majorana Fermi surfaces. This results in the susceptibility maxima occurring at finite momenta (Fig. 3).
To evaluate the susceptibility in angular coordinates, we rewrite the Fermi-velocity of the conduction sea in the momentum basis,
(10) |
where, and . We then parameterize the of the Fermi surface in spherical coordinates
(11) |
in terms of angular variables and and radius . Carrying out the angular integral in equation (9), we numerically find that the susceptibility is maximized (Fig. 3) for:
(12) |
along the principal axes where the magnitude of the wavenumber at the susceptibility maxima (Fig. 3) is proportional to the chemical potential. Here, the cubic symmetry of our model leads to six degenerate wave vectors along which the spinor order prefers to modulate. Such degeneracy is often lifted, as in the case of FFLOLarkin and Ovchinnikov (1964) states, where the order parameter favors amplitude modulation due to attractive interactions between and . This results in ordering of the form , or other superpositions, depending on the level of doping.
Ginzburg Landau Theory. The criteria for amplitude modulation of the pairing is most easily accessible from the Ginzburg Landau theory where the free energy is a function of the magnitude of the spinor order given by,
(13) | ||||
The quadratic degeneracy amongst the order parameters is broken by the quartic coefficients . Depending on the values of , the model exhibits various amplitude-modulated phases. For , the order parameter displays unidirectional modulation and takes the form . In contrast, for , the degeneracy of , , and is lifted by a sixth-order term, resulting in a pair-density wave crystal of the form , arising through the FFLO mechanism Fulde and Ferrell (1964); Larkin and Ovchinnikov (1964).
Conduction electron response. Thus far, we have shown that doping the CPT Kondo lattice model induces amplitude modulation in the electron-Majorana spinor order parameter. This in turn, generates amplitude modulation in the self-energy, , specifically affecting the triplet pairing component of the electronic self-energy:
(14) |
where denotes the normal component of the self-energy, representing the odd-frequency magnetic component of the electrons, while corresponds to the triplet pairing component. As both and are proportional to , they inherit the amplitude modulation of the spinor order parameter. Thus doping the CPT Kondo lattice Coleman et al. (2022) provides a natural mechanism for the spontaneous formation of pair-density waves beyond one dimension, even in the absence of an external magnetic field.
Discussion. In this paper, we have proposed a natural mechanism for the spontaneous formation of pair-density waves, where the underlying superconductivity arises from pairing with a spin liquid. Screening of the spin liquid by the conduction sea then gives rise to triplet pair-density waves. While our model represents a specific case, the underlying concept extends to other forms of unconventional superconductivity believed to result from spin fractionalization, such as the RVB theory of high- superconductivity.
In heavy fermion materials, such pair-density wave (PDW) order has been observed in STM experiments in UTe2 Gu et al. (2023). This PDW order has subsequently been identified as the parent order underlying the charge-density wave (CDW) order through melting transitions Aishwarya et al. (2024). Furthermore, time-reversal symmetry breaking observed in chiral step-edge experiments Jiao et al. (2020) and the onset of re-entrant superconductivity at 40T Aoki et al. (2022), alongside the absence of a two-stage transition and the fully gapped spectrum revealed by transport measurements Suetsugu et al. (2024); Theuss et al. (2024), have fueled debate Jiao et al. (2020); Ishizuka et al. (2019); Xu et al. (2019); Machida (2020); Bae et al. (2021); Hayes et al. (2021); Machida (2021); Nakamine et al. (2021); Thomas et al. (2021); Fujibayashi et al. (2022); Girod et al. (2022); Rosa et al. (2022); Shaffer and Chichinadze (2022); Wei et al. (2022); Hazra and Volkov (2024); Hazra and Coleman (2023); Iguchi et al. (2023); Ishihara et al. (2023); Matsumura et al. (2023); Chang et al. (2024); Suetsugu et al. (2024); Theuss et al. (2024) over the nature of superconducting order of UTe2. This controversy arises because no known two-dimensional representations of superconducting order for UTe2 can fully reconcile these experimental findings.
While our work does not provide a microscopic mechanism for the superconductivity in UTe2, it does open a way out of this dilemma. Superconductivity breaks the U(1) isospsin symmetry associated with charge conservation. In principle, group theory allows for both singly connected representations of the order parameter, exemplified by the charge 2e BCS pairing, but it also allows for double-group representations of the order parameter, with half-integer isospin. Majorana-mediated superconductivity is a first concrete example of this new class of broken-symmetry. Here the electron-Majorana spinor plays the role analogous to the quark in strong interaction physics. By analogy, new families of superconducting phases are now possible, generated by product group representations analogous to the eight-fold way of mesons. This opens up the a zoo of enriched superconducting phases, with a corresponding diversity of new topological superconductors. Recent work has confirmed Zhuang and Coleman (2024) the potential to form topological states out of such fractionalized order. Importantly, double groups introduce Kramers degeneracy, allowing for broken time-reversal in low-symmetry crystalline environments, providing a clue for for the emergence of time-reversal symmetry-breaking, with a single-step transition emerging out of a heavy fermi liquid Panigrahi et al. (2024), observed in UTe2, despite its low orthorhombic Immm crystal symmetry of UTe2.
Acknowledgements.
Acknowledgments: AP and PC would like to thank Tamaghna Hazra for fruitful discussions. This work was supported by Office of Basic Energy Sciences, Material Sciences and Engineering Division, U.S. Department of Energy (DOE) under Contracts No. DE-SC0012704 (AMT) and DE-FG02-99ER45790 (AP and PC ).References
- Agterberg et al. (2020) Daniel F. Agterberg, J.C. Séamus Davis, Stephen D. Edkins, Eduardo Fradkin, Dale J. Van Harlingen, Steven A. Kivelson, Patrick A. Lee, Leo Radzihovsky, John M. Tranquada, and Yuxuan Wang, “The Physics of Pair-Density Waves: Cuprate Superconductors and Beyond,” Annual Review of Condensed Matter Physics 11, 231–270 (2020).
- Fulde and Ferrell (1964) Peter Fulde and Richard A. Ferrell, “Superconductivity in a Strong Spin-Exchange Field,” Physical Review 135, A550–A563 (1964).
- Larkin and Ovchinnikov (1964) A.I. Larkin and Y.N. Ovchinnikov, “Nonuniform state of superconductors,” Zh. Eksp. Teor. Fiz. 47, 1136–1146 (1964).
- Randeria et al. (2016) Mallika T. Randeria, Benjamin E. Feldman, Ilya K. Drozdov, and Ali Yazdani, “Scanning Josephson spectroscopy on the atomic scale,” Physical Review B 93, 161115 (2016).
- Hücker et al. (2011) M. Hücker, M. V. Zimmermann, G. D. Gu, Z. J. Xu, J. S. Wen, Guangyong Xu, H. J. Kang, A. Zheludev, and J. M. Tranquada, “Stripe order in superconducting La 2-x Bax CuO4 ,” Physical Review B 83, 104506 (2011).
- Zhao et al. (2023) He Zhao, Raymond Blackwell, Morgan Thinel, Taketo Handa, Shigeyuki Ishida, Xiaoyang Zhu, Akira Iyo, Hiroshi Eisaki, Abhay N. Pasupathy, and Kazuhiro Fujita, “Smectic pair-density-wave order in EuRbFe4As4,” Nature 618, 940–945 (2023).
- Liu et al. (2023) Yanzhao Liu, Tianheng Wei, Guanyang He, Yi Zhang, Ziqiang Wang, and Jian Wang, “Pair density wave state in a monolayer high-Tc iron-based superconductor,” Nature 618, 934–939 (2023).
- Gu et al. (2023) Qiangqiang Gu, Joseph P. Carroll, Shuqiu Wang, Sheng Ran, Christopher Broyles, Hasan Siddiquee, Nicholas P. Butch, Shanta R. Saha, Johnpierre Paglione, J. C. Séamus Davis, and Xiaolong Liu, “Detection of a pair density wave state in UTe2,” Nature 618, 921–927 (2023).
- Fradkin et al. (2015) Eduardo Fradkin, Steven A. Kivelson, and John M. Tranquada, “Colloquium: Theory of intertwined orders in high temperature superconductors,” Rev. Mod. Phys. 87, 457–482 (2015).
- Aishwarya et al. (2024) Anuva Aishwarya, Julian May-Mann, Avior Almoalem, Sheng Ran, Shanta R. Saha, Johnpierre Paglione, Nicholas P. Butch, Eduardo Fradkin, and Vidya Madhavan, “Melting of the charge density wave by generation of pairs of topological defects in UTe2,” Nature Physics 20, 964–969 (2024).
- Bardeen et al. (1957) J. Bardeen, L. N. Cooper, and J. R. Schrieffer, “Microscopic theory of superconductivity,” Phys. Rev. 106, 162–164 (1957).
- Coleman et al. (2022) Piers Coleman, Aaditya Panigrahi, and Alexei Tsvelik, “Solvable 3D Kondo Lattice Exhibiting Pair Density Wave, Odd-Frequency Pairing, and Order Fractionalization,” Physical Review Letters 129, 177601 (2022).
- Tsvelik and Coleman (2022) Alexei M. Tsvelik and Piers Coleman, “Order fractionalization in a Kitaev-Kondo model,” Physical Review B 106, 125144 (2022).
- Yao and Lee (2011) Hong Yao and Dung-Hai Lee, “Fermionic magnons, non-abelian spinons, and the spin quantum hall effect from an exactly solvable spin- kitaev model with su(2) symmetry,” Phys. Rev. Lett. 107, 087205 (2011).
- Hermanns and Trebst (2014) M. Hermanns and S. Trebst, “Quantum spin liquid with a majorana fermi surface on the three-dimensional hyperoctagon lattice,” Phys. Rev. B 89, 235102 (2014).
- Jiao et al. (2020) Lin Jiao, Sean Howard, Sheng Ran, Zhenyu Wang, Jorge Olivares Rodriguez, Manfred Sigrist, Ziqiang Wang, Nicholas P. Butch, and Vidya Madhavan, “Chiral superconductivity in heavy-fermion metal UTe2,” Nature 579, 523–527 (2020).
- Aoki et al. (2022) D Aoki, J-P Brison, J Flouquet, K Ishida, G Knebel, Y Tokunaga, and Y Yanase, “Unconventional superconductivity in UTe,” Journal of Physics: Condensed Matter 34, 243002 (2022).
- Suetsugu et al. (2024) Shota Suetsugu, Masaki Shimomura, Masashi Kamimura, Tomoya Asaba, Hiroto Asaeda, Yuki Kosuge, Yuki Sekino, Shun Ikemori, Yuichi Kasahara, Yuhki Kohsaka, Minhyea Lee, Youichi Yanase, Hironori Sakai, Petr Opletal, Yoshifumi Tokiwa, Yoshinori Haga, and Yuji Matsuda, “Fully gapped pairing state in spin-triplet superconductor UTe,” Science Advances 10, eadk3772 (2024).
- Theuss et al. (2024) Florian Theuss, Avi Shragai, Gaël Grissonnanche, Ian M. Hayes, Shanta R. Saha, Yun Suk Eo, Alonso Suarez, Tatsuya Shishidou, Nicholas P. Butch, Johnpierre Paglione, and B. J. Ramshaw, “Single-component superconductivity in UTe2 at ambient pressure,” Nature Physics 20, 1124–1130 (2024).
- Ishizuka et al. (2019) Jun Ishizuka, Shuntaro Sumita, Akito Daido, and Youichi Yanase, “Insulator-Metal Transition and Topological Superconductivity in UTe 2 from a First-Principles Calculation,” Physical Review Letters 123, 217001 (2019).
- Xu et al. (2019) Yuanji Xu, Yutao Sheng, and Yi-feng Yang, “Quasi-Two-Dimensional Fermi Surfaces and Unitary Spin-Triplet Pairing in the Heavy Fermion Superconductor UTe 2,” Physical Review Letters 123, 217002 (2019).
- Machida (2020) Kazushige Machida, “Theory of Spin-polarized Superconductors —An Analogue of Superfluid He A-phase—,” Journal of the Physical Society of Japan 89, 033702 (2020).
- Bae et al. (2021) Seokjin Bae, Hyunsoo Kim, Yun Suk Eo, Sheng Ran, I-lin Liu, Wesley T. Fuhrman, Johnpierre Paglione, Nicholas P. Butch, and Steven M. Anlage, “Anomalous normal fluid response in a chiral superconductor UTe2,” Nature Communications 12, 2644 (2021).
- Hayes et al. (2021) I. M. Hayes, D. S. Wei, T. Metz, J. Zhang, Y. S. Eo, S. Ran, S. R. Saha, J. Collini, N. P. Butch, D. F. Agterberg, A. Kapitulnik, and J. Paglione, “Multicomponent superconducting order parameter in UTe,” Science 373, 797–801 (2021).
- Machida (2021) Kazushige Machida, “Nonunitary triplet superconductivity tuned by field-controlled magnetization: URhGe, UCoGe, and UTe 2,” Physical Review B 104, 014514 (2021).
- Nakamine et al. (2021) Genki Nakamine, Katsuki Kinjo, Shunsaku Kitagawa, Kenji Ishida, Yo Tokunaga, Hironori Sakai, Shinsaku Kambe, Ai Nakamura, Yusei Shimizu, Yoshiya Homma, Dexin Li, Fuminori Honda, and Dai Aoki, “Inhomogeneous Superconducting State Probed by Te NMR on UTe,” Journal of the Physical Society of Japan 90, 064709 (2021).
- Thomas et al. (2021) S. M. Thomas, C. Stevens, F. B. Santos, S. S. Fender, E. D. Bauer, F. Ronning, J. D. Thompson, A. Huxley, and P. F. S. Rosa, “Spatially inhomogeneous superconductivity in UTe 2,” Physical Review B 104, 224501 (2021).
- Fujibayashi et al. (2022) Hiroki Fujibayashi, Genki Nakamine, Katsuki Kinjo, Shunsaku Kitagawa, Kenji Ishida, Yo Tokunaga, Hironori Sakai, Shinsaku Kambe, Ai Nakamura, Yusei Shimizu, Yoshiya Homma, Dexin Li, Fuminori Honda, and Dai Aoki, “Superconducting Order Parameter in UTe Determined by Knight Shift Measurement,” Journal of the Physical Society of Japan 91, 043705 (2022).
- Girod et al. (2022) Clément Girod, Callum R. Stevens, Andrew Huxley, Eric D. Bauer, Frederico B. Santos, Joe D. Thompson, Rafael M. Fernandes, Jian-Xin Zhu, Filip Ronning, Priscila F. S. Rosa, and Sean M. Thomas, “Thermodynamic and electrical transport properties of UTe 2 under uniaxial stress,” Physical Review B 106, L121101 (2022).
- Rosa et al. (2022) Priscila F. S. Rosa, Ashley Weiland, Shannon S. Fender, Brian L. Scott, Filip Ronning, Joe D. Thompson, Eric D. Bauer, and Sean M. Thomas, “Single thermodynamic transition at 2 K in superconducting UTe2 single crystals,” Communications Materials 3, 33 (2022).
- Shaffer and Chichinadze (2022) Daniel Shaffer and Dmitry V. Chichinadze, “Chiral superconductivity in UTe 2 via emergent C 4 symmetry and spin-orbit coupling,” Physical Review B 106, 014502 (2022).
- Wei et al. (2022) Di S. Wei, David Saykin, Oliver Y. Miller, Sheng Ran, Shanta R. Saha, Daniel F. Agterberg, Jörg Schmalian, Nicholas P. Butch, Johnpierre Paglione, and Aharon Kapitulnik, “Interplay between magnetism and superconductivity in UTe 2,” Physical Review B 105, 024521 (2022).
- Hazra and Volkov (2024) Tamaghna Hazra and Pavel A. Volkov, “Pair Kondo effect: A mechanism for time-reversal symmetry breaking superconductivity in UTe 2,” Physical Review B 109, 184501 (2024).
- Hazra and Coleman (2023) Tamaghna Hazra and Piers Coleman, “Triplet Pairing Mechanisms from Hund’s-Kondo Models: Applications to UTe 2 and CeRh 2 As 2,” Physical Review Letters 130, 136002 (2023).
- Iguchi et al. (2023) Yusuke Iguchi, Huiyuan Man, S. M. Thomas, Filip Ronning, Priscila F.S. Rosa, and Kathryn A. Moler, “Microscopic Imaging Homogeneous and Single Phase Superfluid Density in UTe2,” Physical Review Letters 130, 196003 (2023).
- Ishihara et al. (2023) Kota Ishihara, Masaki Roppongi, Masayuki Kobayashi, Kumpei Imamura, Yuta Mizukami, Hironori Sakai, Petr Opletal, Yoshifumi Tokiwa, Yoshinori Haga, Kenichiro Hashimoto, and Takasada Shibauchi, “Chiral superconductivity in UTe2 probed by anisotropic low-energy excitations,” Nature Communications 14, 2966 (2023).
- Matsumura et al. (2023) Hiroki Matsumura, Hiroki Fujibayashi, Katsuki Kinjo, Shunsaku Kitagawa, Kenji Ishida, Yo Tokunaga, Hironori Sakai, Shinsaku Kambe, Ai Nakamura, Yusei Shimizu, Yoshiya Homma, Dexin Li, Fuminori Honda, and Dai Aoki, “Large Reduction in the a -axis Knight Shift on UTe with T = 2.1 K,” Journal of the Physical Society of Japan 92, 063701 (2023).
- Chang et al. (2024) Yung-Yeh Chang, Khoe Van Nguyen, Kuang-Lung Chen, Yen-Wen Lu, Chung-Yu Mou, and Chung-Hou Chung, “Correlation-driven topological Kondo superconductors,” Communications Physics 7, 253 (2024).
- Zhuang and Coleman (2024) Zekun Zhuang and Piers Coleman, “Topological superconductivity in a spin-orbit coupled Kondo lattice,” (2024), arXiv:2410.02171 [cond-mat].
- Panigrahi et al. (2024) Aaditya Panigrahi, Alexei Tsvelik, and Piers Coleman, “Breakdown of order fractionalization in the cpt model,” Phys. Rev. B 110, 104520 (2024).
- Coleman et al. (1994) P. Coleman, E. Miranda, and A. Tsvelik, “Odd-frequency pairing in the Kondo lattice,” Physical Review B 49, 8955–8982 (1994).
I Supplementary Material: Electron-Majorana susceptibility
The 3D Kondo lattice model with Yao-Lee spin-orbit interaction exhibits triplet superconductivity Coleman et al. (2022) at half-filling. The analytical tractability of this model, formulated on the hyperoctagon lattice, stems from nesting between the electron-hole and Majorana Fermi surfaces.
The Hamiltonian for the Kondo lattice (CPT) model consists of three components
(15) |
where describes the nearest neighbor hopping of the conduction electrons, describes the Yao-Lee spin-spin interaction and which couples the conduction sea to the Yao-Lee spin liquid via Kondo interaction,
(16) | ||||
-
1.
Yao-Lee spin liquid: The Yao-Lee term involves a Kitaev interaction between the orbitals between nearest neighbors decorated with Heisenberg interaction between the spins . The result of the Kiteav orbital frustration is majorana fractionalization of the orbitals and the spins into . Consequently, the fractionalized representation of the Yao-Lee Hamiltonian is obtained by noting in the physical Hilbert space to be,
(17) Where, similar to Kitaev spin liquid become static gauge fields, (i.e. ).
Noting that Yao-Lee spin liquid undergoes an Ising transition where below Hermanns and Trebst (2014); Coleman et al. (2022); Panigrahi et al. (2024), leading to confinement of visons and deconfinement of Majoranas , we make the gauge choice Hermanns and Trebst (2014) leading to a translationally invariant Yao-Lee Hamiltonian. Transforming to a momentum basis,
(18) where is the position of the unit-cell in the BCC lattice, is the number of primitive unit cells in the lattice and is the site index within each unit cell. The Yao-Lee Hamiltonian is given by,
(19) where (19) described a four-band Hamiltonian for three species of Majorana with a global rotational symmetry, where
(20) Moreover, the momentum spans the half-Brillouin zone. The Yao-Lee spin liquid can be described at low energies by an effective single-band model obtained by projecting to the lowest energy band. The dispersion for the said band is obtained by throwing away higher order terms in the characteristic equation for ,
(21) where, , and s are the eigenvalues of .
The projected band describing the low energy physics of Yao-Lee Hamiltonian then takes the form,
(22) using which, the projected Yao-Lee Hamiltonian is expressed using as follows,
(23) -
2.
Conduction electrons: Similar to the Yao-Lee spin liquid, the conduction sea is also described by a four-band Hamiltonian, with an electron and a hole pocket (2). To illustrate the perfect nesting between the conduction sea and the Majorana spinons, one carries out a non-singular gauge transformation which shifts the Fermi-surface to i.e., the pointColeman et al. (2022). Following said gauge transformation, a four-band Hamiltonian describes the conduction sea,
(24) where is given in equation (20).
The low energy physics of the conduction band can be described by a projected single band , with taking the form (22). Further, the conduction sea can be restricted to (\mancube), by rewriting the Hamiltonian in terms of the Balian-Werthamer spinor as,
(25) Here, coupling the chemical potential with indicates the presence of an electron and a hole Fermi surface in the conduction sea.
-
3.
Kondo interaction: The Majorana fractionalization of spins due to Yao-Lee interaction allows for an analytic mean-field treatment of Kondo interaction without resorting to Gutzwiller projection. In the Majorana representation the spins are expressed as , which allows their rewriting as
(26) The factorized form of the Kondo interaction (26) allows for Hubbard Stratenovich transformation of the Kondo interaction in terms of a spinor order ,
(27) Giving us the mean-field interaction term in equation (27).
A more compact version of the mean-field Kondo interaction is obtained by switching to the Balian Werthamer spinor representation for conduction electrons and the spinor order parameter,
(28)
At half-filling, due to nesting between the electron, hole, and Majorana Fermi surfaces, the model exhibits a logarithmic divergence in the electron-Majorana susceptibility. This leads to the condensation of the uniform spinor order for infinitesimal Kondo coupling, akin to a Peierls instability.
Away from half-filling, this nesting is disrupted by the presence of a non-zero chemical potential, which causes the electron Fermi surface to expand and the hole Fermi surface to contract, while the Majorana Fermi surface remains unaffected. The chemical potential thus acts as a cut-off for the logarithmic instability associated with superconductivity, stabilizing a finite momentum order parameter.

This mechanism bears a strong analogy to singlet superconductors under an external magnetic field, where Zeeman splitting of the spin-polarized Fermi surface leads to the formation of Fulde-Ferrell-Larkin-Ovchinnikov (FFLO) Fulde and Ferrell (1964); Larkin and Ovchinnikov (1964) pair-density waves. In both cases, pair-density waves emerge when the superconducting order parameter develops finite momentum.
The formation of such incommensurate pair density waves can be understood by examining the electron-Majorana susceptibility, which peaks at finite momentum. This susceptibility profile reveals that the ground-state energy of the system is minimized when the superconducting order parameter acquires finite momentum, thereby stabilizing the pair density wave phase.
Using a similar approach to FFLO, we show that the spinor order gains finite momentum in the analytically tractable Kondo lattice model Coleman et al. (2022) as one moves away from half-filling,
(29) |
Here, is the spinor orientation, and is the magnitude of the uniform order parameter with a spatial modulation with momentum .
Consequently, the mean-field Kondo interaction in Balian-Werthamer representation gains a finite momentum exchange,
(30) |
Here, in the exponent signifies that the holes gain opposite momentum to the electrons under Fourier transformation.
Upon carrying out a Fourier transformation, the Kondo interaction in the momentum basis becomes,
(31) |
This compact notation expresses the exchange of and Balian-Werthamer spinor which is diagrammatically depicted in Fig. 4.
In order to show that the finite momentum order parameter is preferred over the zero-momentum order parameter, one computes the susceptibility for spinor ordering. Where the finite momentum spinor order is preferred, the susceptibility is maximum at the said momentum .
Thus, to demonstrate the energetic favorability of the finite momentum order parameter, we begin by computing the static susceptibility (Fig. 4) for electron-majorana condensation,
(32) |
Where, is the orientation of the BW order parameter, is the Green’s function for the Balian-Werthamer spinor at Matsubara frequency and momentum , and is dimensional Identity matrix, and is the isospin matrix for Balian Werthamer representation, with being Pauli matrices. And, is the Green’s function for Majorana with a Matsubara frequency and momentum .
Due to rotational symmetry of Majoranas, the Green’s function being species independent, acts as an identity matrix, allowing us to rewrite the susceptibility as,
(33) |
Further simplification is made by noting that and are projection operatorsColeman et al. (1994, 2022); Tsvelik and Coleman (2022). Moreover, projects out a Majorana component of the conduction sea which remains gapless upon condensation of the fractionalized spinor order. Additionally, introducing the electron-Majorana susceptibility and hole-Majorana susceptibility ,
(34) | ||||
(35) |
We obtain the following expression for the susceptibility to form fractionalized order:
(36) |
Thus, the task of calculating the susceptibility of the fractionalized order reduces to calculating the sum of electron-Majorana susceptibility and hole-Majorana susceptibility . These susceptibilities are calculated by first carrying out the Matsubara summation over to obtain
(37) |
and,
(38) |
Where is the Fermi-Dirac function.
We rewrite the momentum integral along the constant energy contours to compute the spinor-susceptibility. This is done by noting that the density of states of the electrons and Majoranas around the Fermi surface remains constant. Further, since the Fermi-Dirac function becomes a Heavyside-Theta function at zero temperature i.e. , the electron-Majorana susceptibility becomes,
(39) |
Where Coleman et al. (2022) is the density of states and is the bandwidth of the conduction sea. Here, we have carried out a Taylor expansion to obtain the above expression for electron-Majorana susceptibility, with the Fermi-velocity being the gradient of the dispersion at the Fermi surface.
Accounting for the theta functions, the energy integral takes the form:
(40) |
Where the energy integral is carried out analytically to obtain the following expression for ,
(41) |
Taking the limits for the energetic integral followed by a simplification of the above equation yields the following expression for the susceptibility,
(42) |
Where, is the density of states for , and the spinor order susceptibility , as given in equation (36). This susceptibility maximizes for with . These orders are degenerate, and the breaking of these degeneracy by higher order terms in the Ginzburg Landau theory leads to amplitude modulation of the spinor-order, which subsequently leads to amplitude modulation of triplet pairing.
Finally, as the chemical potential cuts off the logarithmic instability, the transition occurs at a finite Kondo coupling . For small chemical potentials, the required Kondo coupling remains sufficiently low due to Stoner’s criterion, . This ensures that the model remains analytically tractable over a range of chemical potentials . Consequently, this model is one of the rare examples where the spontaneous formation of pair-density waves can be reliably established without the need for an external field beyond one-dimension.