Skip to main content
Log in

On the Exact Solution for a Luttinger Liquid with Repulsion and a Single Point Impurity

  • ELECTRONIC PROPERTIES OF SOLID
  • Published:
Journal of Experimental and Theoretical Physics Aims and scope Submit manuscript

Abstract

The expression for the conductance of a 1D channel, which has been obtained using the well-known exact solution, is analyzed. It is shown that in the case of strong electron—electron interaction, the slowest (linear in frequency) asymptotics of the conductance is determined by the behavior of electron—electron interaction in the region of transition from 1D to 3D motion realized near the impurity.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Notes

  1. It was shown in [7] that the renormalized coefficient of reflection for an LL with attraction is obtained from the transmission coefficient for LL with repulsion using the replacement of (\({{{v}}_{c}}\)K0 ↔ \({v}_{c}^{{ - 1}}\); R0). This property is exact and is observed for any type and intensity of the ee interaction. In the proof, we presume only (at lest asymptotic) convergence of the series in perturbation theory.

  2. This statement actually coincides with the initial assumption of the renorm group theory: the observable quantities calculated using unknown exact Hamiltonian converge. This means that in UV (not IR) region, there exists a scale (introduced by regularization in the Gell-Mann–Low approach), beginning with which the deep UV region gives zero contribution to the observables. The entire difference from the “conventional” problem diverging for \({{{v}}_{c}}\) = 1/2 [8] is that for \({{{v}}_{c}}\) = 2, the expression for the conductance calculated using the exact solution with the KF Hamiltonian converges.

  3. The analytic continuation was discussed in detail in [8].

  4. In [5], the total frequency dependence of the conductance (Eq. (5.6)) was obtained, which followed from the exact solution for the KF Hamiltonian and coincided with expression (17).

REFERENCES

  1. C. L. Kane and M. P. A. Fisher, Phys. Rev. B 46, 15233 (1992).

    Article  ADS  Google Scholar 

  2. A. Furusaki and N. Nagaosa, Phys. Rev. B 47, 4631 (1993).

    Article  ADS  Google Scholar 

  3. A. Furusaki, Phys. Rev. B 56, 9352 (1997).

    Article  ADS  Google Scholar 

  4. L. I. Glazman, K. A. Matveev, and D. Yue, Phys. Rev. B 49, 1966 (1994).

    Article  ADS  Google Scholar 

  5. P. Fendley, A. W. W. Ludwig, and H. Saleur, Phys. Rev. B 52, 8934 (1995).

    Article  ADS  Google Scholar 

  6. P. Fendley, A. W. W. Ludwig, and H. Saleur, Phys. Rev. Lett. 74, 3005 (1995).

    Article  ADS  Google Scholar 

  7. V. V. Afonin and V. Yu. Petrov, JETP Lett. 97, 507 (2013).

    Article  ADS  Google Scholar 

  8. V. V. Afonin, J. Exp. Theor. Phys. 136, 207 (2023).

    Article  ADS  Google Scholar 

  9. M. Peskin and D. Schroeder, An Introduction to Quantum Field Theory (Addison-Wesley, Reading, MA, 1996).

    Google Scholar 

  10. S. L. Adler and R. F. Dashen, Current Algebras and Applications to Particle Physics (W. A. Benjamin, New York, 1968).

    MATH  Google Scholar 

  11. V. J. Emery, in Highly Conducting One-Dimensional Solids (Plenum, New York, 1979), p. 327.

    Google Scholar 

  12. L. D. Landau, Sov. Phys. JETP 7, 19 (1937).

    Google Scholar 

  13. V. V. Afonin and V. Yu. Petrov, J. Exp. Theor. Phys. 107, 542 (2008).

    Article  ADS  Google Scholar 

  14. J. M. Kosterlitz and D. J. Thouless, J. Phys. C 6, 1181 (1973).

    Article  ADS  Google Scholar 

  15. V. V. Afonin and V. Yu. Petrov, Found. Phys. 40, 190 (2010).

    Article  ADS  MathSciNet  Google Scholar 

  16. K.-V. Pham, M. Gabay, and P. Lederer, Phys. Rev. B 61, 1637 (2000).

    Article  Google Scholar 

  17. A. F. Andreev, Sov. Phys. JETP 19, 1126 (1964).

    Google Scholar 

  18. P. G. de Gennes, Superconductivity of Metals and Alloys (W. A. Benjamin, New York, 1966).

    MATH  Google Scholar 

  19. V. V. Afonin and V. Yu. Petrov, JETP Lett. 109, 762 (2019).

    Article  ADS  Google Scholar 

  20. J. von Delft and H. Schoeller, cond-mat/9805275v3 (1998).

  21. H. Bateman and A. Erdelyi, Higher Transcendental Functions (McGraw-Hill, New York, 1953), Vol. 2, p. 148.

    MATH  Google Scholar 

Download references

ACKNOWLEDGMENTS

The author thanks Ya.M. Beltyukov and Yu.M. Galperin for reading the manuscript and discussion.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to V. V. Afonin.

Ethics declarations

The author declares that he has no conflicts of interest.

Additional information

Translated by N. Wadhwa

REDUCTION OF THE HAMILTONIAN TO DIAGONAL FORM

REDUCTION OF THE HAMILTONIAN TO DIAGONAL FORM

1.1 A. “Schrödinger Equation” for Rotation Matrix

To calculate correlator \(\mathcal{S}\)(ω), we must pass from the Schrödinger to the Heisenberg representation. This can easily be done for free fermions, for which the differences between one representation from the other is reduced to the emergence of factor exp (–i\({{\epsilon }_{n}}\)t). For this purpose, expression (11) must be diagonalized. It should be noted that variable ζ (now, Φn = 0) appearing in the Hamiltonian describes the variation (and not total number) of chiral pairs in the ground state. It follows hence that the rotation matrix is determined only by the scattering process (i.e., transitions between states Φn ↔ Φn'; n, n' ≠ 0, and Φn ↔ Φ0. Therefore, the matrix diagonalizing expression (11) has form

$$\begin{gathered} \lambda ({{\epsilon }_{n}}) = \frac{1}{L}\sum\limits_{m \ne 0}^{} {S({{\epsilon }_{n}},{{p}_{m}})\Phi ({{p}_{m}}) + Z({{\epsilon }_{n}}){{\Phi }_{0}}} , \\ {{\lambda }_{0}} = \frac{1}{L}\sum\limits_{m \ne 0}^{} {T({{p}_{m}})\Phi ({{p}_{m}}).} \\ \end{gathered} $$
(19)

(The arguments of the elements of this matrix can be written as \({{\epsilon }_{n}}\), pm for not to confuse the indices of rotated and nonrotated fields; in this notation, \({{\epsilon }_{n}}\) and \({{{v}}_{c}}\)pm are the emerges of “rotated” and “nonrotated” fields, respectively.)

The inverse matrix of the transition can be written as

$$\begin{gathered} \Phi ({{p}_{{{{m}_{1}}}}}) = \frac{1}{L}\sum\limits_{{{n}_{1}} \ne 0}^{} {{{S}^{{ - 1}}}({{\epsilon }_{{{{n}_{1}}}}},{{p}_{{{{m}_{1}}}}})\lambda ({{\epsilon }_{{{{n}_{1}}}}}) + {{Z}^{{ - 1}}}({{p}_{{{{m}_{1}}}}})} {{\lambda }_{0}}, \\ {{\Phi }_{0}} = \frac{1}{L}\sum\limits_{{{n}_{1}} \ne 0}^{} {{{T}^{{ - 1}}}({{\epsilon }_{{{{n}_{1}}}}})\lambda ({{\epsilon }_{{{{n}_{1}}}}}).} \\ \end{gathered} $$
(20)

To obtain expressions for the elements of the inverse matrix, we substitute Eqs. (20) into (19):

$$\begin{gathered} \lambda ({{\epsilon }_{n}}) = \frac{1}{L}\sum\limits_{m \ne 0}^{} {S({{\epsilon }_{n}},{{p}_{m}})} \\ \times \,\left[ {\frac{1}{L}\sum\limits_{{{n}_{1}} \ne 0}^{} {{{S}^{{ - 1}}}({{\epsilon }_{{{{n}_{1}}}}},{{p}_{m}})\lambda ({{\epsilon }_{{{{n}_{1}}}}}) + {{Z}^{{ - 1}}}({{\epsilon }_{n}}){{\lambda }_{0}}} } \right] \\ \, + Z({{\epsilon }_{n}})\frac{1}{L}\sum\limits_{{{n}_{1}} \ne 0}^{} {{{T}^{{ - 1}}}({{\epsilon }_{{{{n}_{1}}}}})\lambda ({{\epsilon }_{{{{n}_{1}}}}}).} \\ \end{gathered} $$

By varying this expression with respect to λn, we obtain

$$L = \frac{1}{L}\sum\limits_{m \ne 0}^{} {S({{\epsilon }_{n}},{{p}_{m}}){{S}^{{ - 1}}}({{\epsilon }_{n}},{{p}_{m}}) + Z({{\epsilon }_{n}}){{T}^{{ - 1}}}({{\epsilon }_{n}}).} $$

On the other hand,

$$\{ \lambda ({{\epsilon }_{n}}),\lambda ({{\epsilon }_{m}})\} = L{{\delta }_{{n, - m}}}.$$

Substituting Eq. (19) into this expression, we obtain

$$L = \frac{1}{L}\sum\limits_{m \ne 0}^{} {S({{\epsilon }_{n}},{{p}_{m}})S({{\epsilon }_{{ - n}}},{{p}_{{ - m}}}) + Z({{\epsilon }_{{ - n}}})Z({{\epsilon }_{n}}).} $$
(21)

Comparing these two expressions, we see that

$${{S}^{{ - 1}}}({{\epsilon }_{n}},{{p}_{m}}) = S({{\epsilon }_{{ - n}}},{{p}_{{ - m}}}),\quad {{T}^{{ - 1}}}({{\epsilon }_{n}}) = Z({{\epsilon }_{{ - n}}}).$$
(22)

To obtain the equations for the rotation matrix, we must compare commutator

$${{[\lambda ({{\epsilon }_{n}}),{{\mathcal{H}}_{\lambda }}]}_{ - }} = {{\epsilon }_{n}}\lambda ({{\epsilon }_{n}})$$

with the same commutator written in terms of “nonrotated” fields Φ. Therefore, we must calculate commutator

$$\begin{gathered} \frac{{{{{v}}_{c}}}}{{2L}}\left[ {\sum\limits_{n \ne 0}^{} {{{p}_{n}}{{{\hat {\Phi }}}_{{ - n}}}{{{\hat {\Phi }}}_{n}}} + 2i{{\gamma }_{2}}} \right.{{\Phi }_{0}}\frac{1}{L}\sum\limits_{n \ne 0}^{} {{{{\hat {\Phi }}}_{n}}} , \\ {{\left. {\frac{1}{L}\sum\limits_{m \ne 0}^{} {S({{\epsilon }_{n}},{{p}_{m}})\Phi ({{p}_{m}}) + Z({{\epsilon }_{n}}){{\Phi }_{0}}} } \right]}_{ - }}. \\ \end{gathered} $$

The term containing the kinetic energy gives

$$ - \frac{{{{{v}}_{c}}}}{L}\sum\limits_{{{m}_{1}} \ne 0}^{} {S({{\epsilon }_{n}},{{p}_{{{{m}_{1}}}}}){{p}_{{{{m}_{1}}}}}\Phi ({{p}_{{{{m}_{1}}}}}).} $$

It remains for us to consider commutator

$$\begin{gathered} \left[ {2i{{\gamma }_{2}}{{\Phi }_{0}}\frac{1}{L}\sum\limits_{n \ne 0}^{} {{{{\hat {\Phi }}}_{n}},} } \right. \\ {{\left. {Z({{\epsilon }_{n}}){{\Phi }_{0}} + \frac{1}{L}\sum\limits_{{{m}_{1}} \ne 0}^{} {S({{\epsilon }_{n}},{{p}_{{{{m}_{1}}}}})\Phi ({{p}_{{{{m}_{1}}}}})} } \right]}_{ - }}. \\ \end{gathered} $$

The first term gives

$$ - i{{\gamma }_{2}}Z({{\epsilon }_{n}})\frac{1}{L}\sum\limits_{n \ne 0}^{} {{{{\hat {\Phi }}}_{n}},} $$

while the second term yields

$$2i{{\gamma }_{2}}{{\Phi }_{0}}\frac{1}{L}\sum\limits_{{{m}_{1}} \ne 0}^{} {S({{\epsilon }_{n}},{{p}_{{{{m}_{1}}}}}).} $$

Ultimately, we obtain equality

$$\begin{gathered} - \frac{{{{{v}}_{c}}}}{L}\sum\limits_{{{m}_{1}} \ne 0}^{} {S({{\epsilon }_{n}},{{p}_{{{{m}_{1}}}}}){{p}_{{{{m}_{1}}}}}\Phi ({{p}_{{{{m}_{1}}}}})} \\ + 2i{{\gamma }_{2}}{{\Phi }_{0}}\frac{1}{L}\sum\limits_{{{m}_{1}} \ne 0}^{} {S({{\epsilon }_{n}},{{p}_{{{{m}_{1}}}}}) - i{{\gamma }_{2}}Z({{\epsilon }_{n}})\frac{1}{L}\sum\limits_{n \ne 0}^{} {{{{\hat {\Phi }}}_{n}}} } \\ = - {{\epsilon }_{n}}\left( {\frac{1}{L}\sum\limits_{{{m}_{1}} \ne 0}^{} {S({{\epsilon }_{n}},{{p}_{{{{m}_{1}}}}}){{p}_{{m1}}}\Phi ({{p}_{{{{m}_{1}}}}}) + Z({{\epsilon }_{n}}){{\Phi }_{0}}} } \right). \\ \end{gathered} $$

This relation must hold for any fields Φ; therefore,

$$\begin{gathered} ({{\epsilon }_{n}} - {{{v}}_{c}}{{p}_{m}})S({{\epsilon }_{n}},{{p}_{m}}) = i{{\gamma }_{2}}Z({{\epsilon }_{n}}), \\ {{\epsilon }_{n}}Z({{\epsilon }_{n}}) = - 2i{{\gamma }_{2}}\frac{1}{L}\sum\limits_{{{m}_{1}} \ne 0}^{} {S({{\epsilon }_{n}},{{p}_{{{{m}_{1}}}}})} . \\ \end{gathered} $$
(23)

Thus, we have obtained the following equation determining the spectrum:

$${{\epsilon }_{n}} = 2\gamma _{2}^{2}\frac{1}{L}\sum\limits_{m \ne 0}^{} {\frac{1}{{{{\epsilon }_{n}} - {{{v}}_{c}}{{p}_{m}}}}} .$$
(24)

Passing to dimensionless energy yn = L\({{\epsilon }_{n}}\)/2π\({{{v}}_{c}}\), we get

$${{y}_{n}} = {{d}^{2}}\left[ {\pi {\text{cot}}(\pi {{y}_{n}}) - \frac{1}{{{{y}_{n}}}}} \right],\quad {{d}^{2}}\, = \,2L{{\left( {\frac{{{{\gamma }_{2}}}}{{2\pi {{{v}}_{c}}}}} \right)}^{2}} \gg 1,$$
(25)

or

$$\tan (\pi {{y}_{n}}) = \frac{{\pi {{d}^{2}}{{y}_{n}}}}{{{{d}^{2}} + y_{n}^{2}}}.$$

It should be noted that [γ2] ~ \(\sqrt {\mathcal{M}} \); therefore, the thermodynamic transition in our case indicates the disregard corrections in 1/\(\mathcal{M}L\) for \(\mathcal{M}L\) \( \gg \) 1, but not requires the passage to the limit L → ∞. In the zeroth approximation in this parameter, we seek the solution in form

$${{y}_{n}} = n + \delta {{y}_{k}},\quad n \gg \delta {{y}_{n}}.$$

Then the shift around each point yn = n satisfies relation

$$\delta {{y}_{n}} = \frac{1}{\pi }\arctan \left( {\frac{{\pi n}}{{1 + {{{(n{\text{/}}d)}}^{2}}}}} \right),$$

where arctan is defined on interval (–π/2, π/2).

Ultimately, the spectrum is given by

$${{\epsilon }_{n}} = \frac{{2\pi {{{v}}_{c}}}}{L}\left( {n + \frac{1}{\pi }\arctan \left( {\frac{{\pi n}}{{1 + {{{(n{\text{/}}d)}}^{2}}}}} \right)} \right)$$
(26)

for n \( \gg \) δyn. However, for n \( \gg \) 1, this condition is always satisfied because arctangent is defined on interval (–π/2, π/2), i.e., ~1. It follows hence that the level shift is small and can be ignored. We have the spectrum satisfying periodic boundary conditions: yn = n. A transition to the thermodynamic limit is performed conventionally, \(\mathcal{M}L\) \( \gg \) 1, and 2πn/L = k is independent of L. Ultimately, the spectrum of the fields that are not scattered by the impurity is also equal to \(\epsilon \)(k) = \({{{v}}_{c}}\)k.

1.2 B. Calculation of Rotation Matrix

Relations (20) and (22) make it possible to express the Guinier field in terms of the fields λn diagonalizing the Hamiltonian, while the normalization condition written in form (21) with account for the first equation of system (23) makes it possible to calculate the required matrix element of rotation matrix Z(\({{\epsilon }_{n}}\)):

$${{\left( {\frac{{{{\gamma }_{2}}}}{{2\pi {{{v}}_{c}}}}} \right)}^{2}}L\sum\limits_{{{m}_{1}} \ne 0}^{} {\frac{{Z({{\epsilon }_{n}})Z( - {{\epsilon }_{n}})}}{{{{{({{y}_{n}} - {{m}_{1}})}}^{2}}}}} + Z({{y}_{n}})Z( - {{y}_{m}}) = L.$$
(27)

To evaluate the sum on the left-hand side of this expression, we note that

$$\begin{gathered} - {{\partial }_{y}}\sum\limits_{{{m}_{1}} \ne 0}^{} {\frac{1}{{{{y}_{n}} - {{m}_{1}}}}} = - {{\partial }_{y}}\left( {\pi \cot (\pi y) - \frac{1}{y}} \right) \\ = {{\pi }^{2}}\left( {1 + \frac{{1 - {{{\sin }}^{2}}(\pi y)}}{{{{{\sin }}^{2}}(\pi y)}}} \right) - \frac{1}{{{{y}^{2}}}} \\ = {{\pi }^{2}} + {{\pi }^{2}}{{\cot }^{2}}(\pi y) - \frac{1}{{{{y}^{2}}}} = {{\pi }^{2}} + \frac{2}{{{{d}^{2}}}} + \frac{{y_{k}^{2}}}{{{{d}^{4}}}}. \\ \end{gathered} $$

Disregarding corrections in 1/L, we obtain the final expression for the transition of matrix elements to the diagonal form:

$$Z({{\epsilon }_{n}}) = \frac{{2\sqrt 2 i{{\gamma }_{2}}{\text{sgn}}({{\epsilon }_{n}})}}{{\sqrt {{{\mu }^{2}} + \epsilon _{n}^{2}} }},$$
(28)
$$S({{\epsilon }_{n}},{{p}_{m}}) = - \frac{{2\sqrt 2 \gamma _{2}^{2}{\text{sgn}}({{\epsilon }_{n}})}}{{\sqrt {{{\mu }^{2}} + \epsilon _{n}^{2}} ({{\epsilon }_{n}} - {{{v}}_{c}}{{p}_{m}})}},$$
(29)

where μ = \(\gamma _{2}^{2}\)/\({{{v}}_{c}}\) ~ \(\mathcal{M}\). This allows us to express the Guinier field in the Heisenberg representation in terms of fields λ(\({{\epsilon }_{n}}\)) of free quasiparticles:

$$\zeta (t) = - \frac{{4i{{\gamma }_{2}}}}{L}\sum\limits_{n \ne 0}^{} {\frac{{{\text{sgn}}({{\epsilon }_{n}})}}{{\sqrt {{{\mu }^{2}} + \epsilon _{n}^{2}} }}\lambda ({{\epsilon }_{n}})\exp ( - i{{\epsilon }_{n}}t).} $$
(30)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Afonin, V.V., Petrov, V.Y. On the Exact Solution for a Luttinger Liquid with Repulsion and a Single Point Impurity. J. Exp. Theor. Phys. 137, 384–394 (2023). https://doi.org/10.1134/S1063776123090017

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S1063776123090017

Navigation