Skip to main content

Indirect Exchange Interaction Mediated by Dirac Fermions

  • Chapter
  • First Online:
Magnetism in Topological Insulators
  • 952 Accesses

Abstract

Based on the model of s-d interaction discussed in Chap. 6, the indirect exchange interaction between a pair of magnetic impurities is calculated. The interaction carries the signature of the topological surface states. The chapter discusses possible magnetic textures induced by the indirect exchange.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 79.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 99.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 129.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. Zhu, J.-J., Yao, D.X., Zhang, S.-C., Chang, K.: Phys. Rev. Lett. 106, 097201 (2011)

    Article  Google Scholar 

  2. Bloembergen, N., Rowland, T.J.: Phys. Rev. 97, 1679 (1955)

    Article  Google Scholar 

  3. Abrikosov, A.A., Gorkov, L.P., Dzyaloshinski, I.E.: Quantum Field Theoretical Methods in Statistical Mechanics. Pergamon, New York (1965)

    MATH  Google Scholar 

  4. Ruderman, M.A., Kittel, C.: Phys. Rev. 96, 99 (1954)

    Article  Google Scholar 

  5. Kasuya, T.: Prog. Theor. Phys. 16, 45 (1956)

    Article  Google Scholar 

  6. Yosida, K.: Phys. Rev. 106, 893 (1957)

    Article  Google Scholar 

  7. De Gennes, P.G.: J. Phys. Radium. 23, 630 (1962)

    Article  Google Scholar 

  8. Mattis, D.C.: The Theory of Magnetism Made Simple. World Scientific, Singapore (2006)

    Book  Google Scholar 

  9. Bloembergen, N., Rowland, T.J.: Phys. Rev. 97(6), 1680 (1955)

    Article  Google Scholar 

  10. Litvinov, V.I.: Sov. Phys. Semicond. 19, 345 (1985)

    Google Scholar 

  11. Abrikosov, A.A.: J. Low Temp. Phys. 39(1/2), 217 (1980)

    Article  Google Scholar 

  12. Liu, L., Bastard, G.: Phys. Rev. B25(1), 487 (1982)

    Article  Google Scholar 

  13. Abanin, D.A., Pesin, D.A.: Phys. Rev. Lett. 106, 136802 (2011)

    Article  Google Scholar 

  14. Fischer, B., Klein, M.W.: Phys. Rev. B11, 2025–2029 (1975)

    Article  Google Scholar 

  15. Litvinov, V.I., Dugaev, V.K.: Phys. Rev. B58(7), 3584 (1996)

    Article  Google Scholar 

  16. Lu, H.-Z., Shan, W.-Y., Wang, Y., Niu, Q., Shen, S.-Q.: Phys. Rev. 115407, B81 (2010)

    Google Scholar 

  17. Shan, W.-Y., Lu, H.-Z., Shen, S.-Q.: New J. Phys. 12, 043048 (2010)

    Article  Google Scholar 

  18. Shen, S.-Q., Shan, W.-Y., Lu, H.-Z.: arXiv:1009.5502v3 [cond-mat.mes-hall]

    Google Scholar 

  19. Litvinov, V.I., Dugaev, V.K.: Phys. Rev. Lett. 86(24), 5593 (2001)

    Article  Google Scholar 

  20. Dugaev, V.K., Litvinov, V.I.: Superlattices Microstruct. 16, 413 (1994)

    Article  Google Scholar 

  21. Litvinov, V.I., Dugaev, V.K.: Phys. Rev. B. 58, 3584 (1998)

    Article  Google Scholar 

  22. Dugaev, V.K., Litvinov, V.I., Barnas, J.: Phys. Rev. B. 74, 224438 (2006)

    Article  Google Scholar 

  23. Litvinov, V.: Wide Bandgap Semiconductor Spintronics. Pan Stanford, Singapore (2016)

    Book  Google Scholar 

  24. Liu, Q., Liu, C.-X., Xu, C., Qi, X.-L., Zhang, S.-C.: Phys. Rev. Lett. 102, 156603 (2009)

    Article  Google Scholar 

  25. Efimkin, D.K., Galitski, V.: Phys. Rev. B. 89, 115431 (2014)

    Article  Google Scholar 

  26. Hsieh, T.H., Lin, H., Liu, J., Duan, W., Bansil, A., Fu, L.: Nat. Commun. 3, 982 (2012)

    Article  Google Scholar 

  27. Tanaka, Y., Ren, Z., Sato, T., Nakayama, K., Souma, S., Takahashi, T., Segawa, K., Ando, Y.: Nat. Phys. 8, 800 (2012)

    Article  Google Scholar 

  28. Dugaev, V.K., Litvinov, V.I., Barnas, J.: Phys. Rev. B74, 224438 (2006)

    Article  Google Scholar 

  29. Litvinov, V.I.: Sov. Phys. Solid State. 27(4), 740 (1985)

    Google Scholar 

  30. Abramovitz, M., Stegun, I.A. (eds.): Handbook of Mathematical Functions. National Bureau of Standards, Gaithersburg (1964)

    Google Scholar 

  31. Litvinov, V.I.: Phys. Rev. B89, 235316 (2014)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Appendices

Appendix 1. Spin Texture

The spin configuration, governed by the terms t1 = C1(S1 × R)z(S2 × R)z and t2 = C2(S1R)(S2R) in pair interaction (7.30) and (7.32), stems from the minimum energy it delivers when one rotates spins in the pair. If one rotates S1 and S2 in the geometry of Fig. 7.4, the interaction energy corresponding to t1 and t2, depends on angles γ1 and γ1 as shown in Fig. 7.15.

Fig. 7.15
figure 15

t1 (a) and t2 (b) energy surfaces in (γ1, γ2) plane, C1, 2 > 0. Insets show the spin alignment corresponding to the minimum energy

The terms, t1 and t2, favor different spin configurations with respect to the inter-spin vector (see insets in Fig. 7.15a, b) while they are similar in their antiferromagnetic nature as long as C1,2 > 0. If any of the coefficients C1,2 becomes negative, the corresponding ordering turns ferromagnetic.

The actual spin texture in a magnetically doped TI film is determined by the total indirect exchange interaction that simultaneously includes competing t1 and t2 contributions as well as the Ising term in (7.30). In addition, the coefficients which are the range functions may be variable in sign. If one considers the magnetic ordering on a small inter-spin distance where sign variations are not relevant, the relative magnitude of the interaction on the surfaces and in the middle of the film make us conclude that surfaces experience out-of-plane ferromagnetism, while the spin glass forms in the inner part of the film.

Appendix 2. Calculation of the RKKY in a Vertically Biased Topological Insulator

In order to calculate the range function (7.33) analytically, we use the conduction band energy spectrum (7.21) expanded up to the k2 term (reference energy is in the middle of the surface energy gap):

$$ {\displaystyle \begin{array}{c}{E}_{\mathrm{c},\uparrow \downarrow }(k)=\frac{1}{2}\sqrt{\varDelta_S^2+4{\tilde{V}}_{as}^2} + \tilde{B}{k}^2\pm {\alpha}_Rk,\\ {}\tilde{B}=D-{Bg}_1\left({g}_1^2+4{V}^2\right)+\frac{g_1^2\ {\tilde{\mathrm{A}}}_2^2}{\sqrt{\varDelta_S^2+4{\tilde{V}}_{as}^2}}\equiv \frac{{\mathrm{\hslash}}^2}{2m},\\ {}{\alpha}_{\mathrm{R}}=2{\tilde{A}}_2V,\kern1em {g}_1=\frac{\varDelta_S}{\sqrt{\varDelta_S^2+4{\tilde{V}}_{as}^2}},\kern1em V=\frac{{\tilde{V}}_{as}}{\sqrt{\varDelta_S^2+4{\tilde{V}}_{as}^2}}.\end{array}} $$
(7.42)

In what follows we suppose that \( \tilde{B} \) (inverse effective mass) is positive and then the spectrum as a function of momentum comprises two parabolas shifted with respect to each other by Rashba spin-splitting (αR ≠ 0).

We use an analytical continuation of Bessel functions to a negative k-half-plane,

$$ {\displaystyle \begin{array}{c}{J}_0\left({kR}_{\Big\Vert}\right)=\frac{1}{2}\left({H}_0^{(1)}\left({kR}_{\Big\Vert}\right)+{H}_0^{(2)}\left({kR}_{\Big\Vert}\right)\right),\\ {}{H}_0^{(1)}\left({ze}^{i\pi}\right)=-{H}_0^{(2)}(z),\end{array}} $$
(7.43)

where \( {H}_0^{\left(1,2\right)}\left({R}_{\Big\Vert }k\right) \) are Hankel functions [30]. The integral over momentum in (7.33) can be represented as an integral along the line shifted in the upper k-half-plane to avoid the branch cut of the function \( {H}_0^{(1)} \) on the negative k-axis:

$$ {I}_{c\uparrow}\left({\omega}_n\right)=\frac{1}{2}{\int}_{-\infty + i\delta}^{\infty + i\delta}\frac{kH_0^{(1)}\left({kR}_{\parallel}\right) dk}{i{\omega}_n-\tilde{B}{k}^2-{\alpha}_{\mathrm{R}}k+{\varepsilon}_{\mathrm{F}}},\kern1em {\varepsilon}_{\mathrm{F}}=\mu -\frac{1}{2}\sqrt{\varDelta_S^2+4{\tilde{V}}_{as}^2\ }. $$
(7.44)

Expression (7.44) accounts for the spin-up band only. To keep the integration over this band, we have to flip αR →  − αR in the course of transition to a negative k half-axis. Because of Kramers degeneracy, mere reflection k →  − k would correspond to transition to the spin-down band.

The Hankel function \( {H}_0^{(1)}(z) \) is analytical in the upper k-half-plane and tends to zero on the arc z → ∞. So, in order to calculate (7.44), we close the integration path in the upper half-plane. The poles of the integrand are

$$ {k}_{1,2}=\frac{1}{2\tilde{B}}\left[-{\alpha}_{\mathrm{R}}\pm \sqrt{\alpha_{\mathrm{R}}^2+4\tilde{B}\left(i{\omega}_n+{\varepsilon}_{\mathrm{F}}\right)}\right]. $$
(7.45)

At this point, it is important to make sure that the poles of the integrand are located on both sides of the real k-axis. Otherwise, the integral would be zero as the path can be closed around the region, where the integrand is an analytical function. Moreover, as we will use (7.44) in a subsequent summation over the Matsubara frequency ωn, it is important to choose branches of poles to hold their positions on the complex k-plane for ωn > 0 and ωn < 0, as shown in Fig. 7.16.

Fig. 7.16
figure 16

Poles of the integrand in (7.44) on the complex k-plane

Finally, the integral (7.44) is found by calculating the residue

$$ {\displaystyle \begin{array}{c}{I}_{c\uparrow}\left({\omega}_n\right)=-\frac{\pi i}{2\tilde{B}}\cdot \frac{-P+\sqrt{P^2+ ix+F}}{\sqrt{P^2+ ix+F}}{H}_0^{(1)}\left[{R}_{\parallel}\left(-P+\sqrt{P^2+ ix+F}\right)\right],\\ {}P=\frac{\alpha_{\mathrm{R}}}{2\tilde{B}};\kern1em F=\frac{\varepsilon_{\mathrm{F}}}{2\tilde{B}};\kern1em x=\frac{\omega_n}{2\tilde{B}}.\end{array}} $$
(7.46)

The integral Ic(ωn), calculated in a similar way, differs from (7.46) by the sign of coefficient P. The combination that enters the range function (7.33) takes the form:

$$ {\displaystyle \begin{array}{l}{I}_{c\uparrow}\left({\omega}_n\right){I}_{c\downarrow}\left({\omega}_n\right)=-\frac{\pi^2}{4{\tilde{B}}^2}\times \left(1-\frac{P^2}{P^2+ ix+F}\right){H}_0^{(1)}\left[{R}_{\parallel}\left(-P+\sqrt{P^2+ ix+F}\right)\right]\\ {}\times {H}_0^{(1)}\left[{R}_{\parallel}\left(P+\sqrt{P^2+ ix+F}\right)\right],\end{array}} $$
(7.47)

The range function contains the sum over Matsubara frequencies which at T → 0 can be replaced with the integral:

$$ {\displaystyle \begin{array}{c}{S}_{\mathrm{M}}=T\sum \limits_{\omega_n}{I}_{c\uparrow}\left({\omega}_n\right){I}_{c\downarrow}\left({\omega}_n\right)=\frac{1}{2\pi i}{\int}_{\varGamma }{I}_{c\uparrow}\left(\upomega \right){I}_{c\downarrow}\left(\upomega \right) d\omega, \\ {}i{\omega}_n\to \omega +i\ \mathit{\operatorname{sign}}\left(\omega \right).\end{array}} $$
(7.48)

Contour Γ = Γ1 + Γ2, shown in Fig. 7.17, avoids the branch cut point x0 =  − R2 − F and the branch cut line determined by the square root in (7.47).

Fig. 7.17
figure 17

Integration contour in (7.48)

The integral over the circle around the pole tends to zero when the radius decreases, and SM can be expressed via integration over the upper and lower sides of path Γ2:

$$ {\displaystyle \begin{array}{l}{S}_{\mathrm{M}}=\frac{i\pi}{4\tilde{B}}{\int}_0^{\sqrt{P^2+F}} ydy\left(1-\frac{P^2}{y^2}\right)\Big\{{H}_0^{(1)}\left[{R}_{\parallel}\left(-P+y\right)\right]{H}_0^{(1)}\left[{R}_{\parallel}\left(P+y\right)\right]\\ {}\kern2.2em -{H}_0^{(1)}\left[{R}_{\parallel}\left(-P-y\right)\right]{H}_0^{(1)}\left[{R}_{\parallel}\left(P-y\right)\right]\Big\},\\ {}\kern6em y=\sqrt{P^2+x+F}.\end{array}} $$
(7.49)

The Hankel function \( {H}_0^{(1)} \) has a branch cut on the negative part of the real axis. To avoid the integration of \( {H}_0^{(1)} \) in this region, we use the relation \( {H}_0^{(1)}\left(-y\right)=-{H}_0^{(2)}(y) \), so (7.49) takes the form

$$ {\displaystyle \begin{array}{c}{S}_{\mathrm{M}}=\frac{i\pi}{4\tilde{B}}\times \left\{{\int}_0^P{yh}_1(y)\left(1-\frac{P^2}{y^2}\right) dy+{\int}_P^{\sqrt{P^2+F}}{yh}_2(y)\left(1-\frac{P^2}{y^2}\right)\right\} dy,\\ {}{h}_1(y)={H}_0^{(2)}\left[{R}_{\parallel}\left(P+y\right)\right]{H}_0^{(1)}\left[{R}_{\parallel}\left(P-y\right)\right]-{H}_0^{(2)}\left[{R}_{\parallel}\left(P-y\right)\right]{H}_0^{(1)}\left[{R}_{\parallel}\left(P+y\right)\right],\\ {}{h}_2(y)={H}_0^{(1)}\left[{R}_{\parallel}\left(y-P\right)\right]{H}_0^{(1)}\left[{R}_{\parallel}\left(y+P\right)\right]-{H}_0^{(2)}\left[{R}_{\parallel}\left(y+P\right)\right]{H}_0^{(2)}\left[{R}_{\parallel}\left(y-P\right)\right].\end{array}} $$
(7.50)

Changing the variable to s = yR and using notations (7.46), we get the final result for the range function (7.34) [31]. In (7.34), we used the identities relating to the Hankel and Newman functions: \( {H}_0^{(1)}(s)={J}_0(s)+{iN}_0(s),{H}_0^{(2)}(s)={J}_0(s)-{iN}_0(s). \)

Appendix 3. Calculation of Bloembergen-Rowland Coupling

Below we illustrate the calculation method taking the integral I2 (7.38) as an example. After integration over angles, we come to the expression

$$ \mathrm{Int}=\sum \limits_{\omega_n}\iint dkdq\frac{kq\ {J}_0\left({kR}_{\Big\Vert}\right){J}_0\left({qR}_{\Big\Vert}\right){\left(i{\omega}_n\right)}^2}{\left({\varepsilon}_k^2+{\tilde{A}}_2^2{k}^2-{\omega}_n^2\right)\left({\varepsilon}_q^2+{\tilde{A}}_2^2{q}^2-{\omega}_n^2\right)}\equiv \sum \limits_{\omega_n}{\left[Q\left({\omega}_n\right)\right]}^2, $$
(7.51)

where J0 is the Bessel function and

$$ {\displaystyle \begin{array}{l}Q\left({\omega}_n\right)={\int}_0^{\infty } kdk\ \frac{\ i{\omega}_n\kern0.50em {J}_0\left({R}_{\parallel }\ k\right)}{{\left(i{\omega}_n\right)}^2-{B}^2{\left({k}^2-{k_0}^2\right)}^2-{\tilde{A}}_2^2{k}^2}\\ {}=\frac{1}{2}{\int}_0^{\infty } kdk\ \frac{\ i{\omega}_n\ }{{\left(i{\omega}_n\right)}^2-{B}^2{\left({k}^2-{k_0}^2\right)}^2-{\tilde{A}}_2^2{k}^2}\left[{H}_0^{(1)}\left({R}_{\Big\Vert }\ k\right)+{H}_0^{(2)}\left({R}_{\Big\Vert }\ k\right)\right].\end{array}} $$
(7.52)

Using the relation \( {H}_0^{(2)}\left({R}_{\Big\Vert }\ k\right)=-{H}_0^{(1)}\left({R}_{\Big\Vert }\ {ke}^{i\pi}\right) \), the second term in (7.52) can be represented as

$$ {\int}_{-\infty + i\delta}^0 ydy\ \frac{\ i{\omega}_n\ {H}_0^{(1)}\left({R}_{\Big\Vert }\ y\right)}{{\left(i{\omega}_n\right)}^2-{B}^2{\left({y}^2-{k_0}^2\right)}^2-{\tilde{A}}_2^2{y}^2}. $$
(7.53)

Combining both terms in (7.52), we obtain

$$ Q(x)=-\frac{1}{2{B}^2}{\int}_{-\infty +i\infty}^{\infty +i\infty } kdk\ \frac{\kern0.5em x\ {H}_0^{(1)}\left({R}_{\Big\Vert }\ k\right)}{\left(k-{k}_1\right)\left(k-{k}_2\right)\left(k-{k}_3\right)\left(k-{k}_4\right)}, $$
(7.54)

where

$$ {k}_{1,2}=\pm {\left[{k}_{\mathrm{M}}^2+\frac{1}{B}\sqrt{x^2-{\varDelta}_S^2/4}\right]}^{1/2},\kern0.5em {k}_{3,4}=\pm {\left[{k}_{\mathrm{M}}^2-\frac{1}{B}\sqrt{x^2-{\varDelta}_S^2/4}\right]}^{1/2},\kern0.5em x=i{\omega}_n. $$
(7.55)

Calculating the integral (7.54), one obtains

$$ {\displaystyle \begin{array}{l}Q(x)=\kern1em \frac{\pi ix\ \left[{H}_0^{(1)}\left({R}_{\Big\Vert }\ {\left[{k}_{\mathrm{M}}^2+\frac{1}{B}\sqrt{x^2-{\varDelta}_S^2/4}\right]}^{1/2}\right)-{H}_0^{(1)}\left(-{R}_{\Big\Vert }{\left[{k}_{\mathrm{M}}^2-\frac{1}{B}\sqrt{x^2-{\varDelta}_S^2/4}\right]}^{1/2}\right)\right]}{4B\sqrt{x^2-{\varDelta}_S^2/4}}\\ \kern2.5em=-\frac{\pi ix\ \left[{H}_0^{(1)}\left({R}_{\Big\Vert }\ {\left[{k}_{\mathrm{M}}^2+\frac{1}{B}\sqrt{x^2-{\varDelta}_S^2/4}\right]}^{1/2}\right)+{H}_0^{(2)}\left({R}_{\Big\Vert }{\left[{k}_M^2-\frac{1}{B}\sqrt{x^2-{\varDelta}_S^2/4}\right]}^{1/2}\right)\right]}{4B\sqrt{x^2-{\varDelta}_S^2/4}}.\end{array}} $$
(7.56)

At T = 0, discrete summation is replaced with integration over x : n → x + i δ sign (x), so the expression (7.56), being used in (7.51), gives

$$ T\sum \limits_{\omega_n}{\left[Q\left({\omega}_n\right)\right]}^2=\frac{1}{2\pi i}{\int}_{\varGamma }{\left[Q(x)\right]}^2 dx. $$
(7.57)

The integration path in (7.57) is shown in Fig. 7.18.

Fig. 7.18
figure 18

Integration path on x-plane in (7.57)

One can deform the contour Γ rotating it by π/2 as the turn goes in the region, where the integrand is an analytical function:

$$ {\displaystyle \begin{array}{c}T\sum \limits_{\omega_n}{\left[Q\left({\omega}_n\right)\right]}^2=\frac{\pi i}{16{B}^2}{\int}_0^{i\infty}\frac{x^2}{x^2-{\varDelta}_S^2/4}\Big[{H}_0^{(1)}\left({R}_{\Big\Vert }\ {\left[{k}_{\mathrm{M}}^2+\frac{1}{B}\sqrt{x^2-{\varDelta}_S^2/4}\right]}^{1/2}\right)\\ {}\kern1em +{H}_0^{(2)}\left({R}_{\Big\Vert }{\left[{k}_{\mathrm{M}}^2-\frac{1}{B}\sqrt{x^2-{\varDelta}_S^2/4}\right]}^{1/2}\right)\Big]{}^2 dx.\end{array}} $$
(7.58)

Changing the variable to \( =\sqrt{1-4{x}^2/{\varDelta}_S^2} \), we express (7.58) as

$$ {\displaystyle \begin{array}{c}T\sum \limits_{\omega_n}{\left[Q\left({\omega}_n\right)\right]}^2=-\frac{\pi {\varDelta}_S}{32{B}^2}{\int}_1^{\infty}\frac{\sqrt{z^2-1}}{\mathrm{z}}\Big[{H}_0^{(1)}\left(\frac{R_{\Big\Vert }}{R_0}\ {\left[{R}_0^2{k}_{\mathrm{M}}^2+ iz\right]}^{1/2}\right)\\ {}\kern0.9em +{H}_0^{(2)}\left(\frac{R_{\Big\Vert }}{R_0}\ {\left[{R}_0^2{k}_{\mathrm{M}}^2- iz\right]}^{1/2}\right)\Big]{}^2 dz\\ {}=-\frac{\pi {\varDelta}_S}{16{B}^2}\underset{1}{\overset{\infty }{\int }}\frac{\sqrt{z^2-1}}{\mathrm{z}}{\left\{\operatorname{Re}\left[{H}_0^{(1)}\left(\frac{R_{\Big\Vert }}{R_0}\ {\left[{R}_0^2{k}_{\mathrm{M}}^2+ iz\right]}^{1/2}\right)\right]\right\}}^2 dz.\end{array}} $$
(7.59)

The expression (7.59) enters the final result (7.39). The terms I1 and I3 in (7.38) were calculated in a similar way.

Rights and permissions

Reprints and permissions

Copyright information

© 2020 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Litvinov, V. (2020). Indirect Exchange Interaction Mediated by Dirac Fermions. In: Magnetism in Topological Insulators. Springer, Cham. https://doi.org/10.1007/978-3-030-12053-5_7

Download citation

  • DOI: https://doi.org/10.1007/978-3-030-12053-5_7

  • Published:

  • Publisher Name: Springer, Cham

  • Print ISBN: 978-3-030-12052-8

  • Online ISBN: 978-3-030-12053-5

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics