Next Article in Journal
Dependency of the Charge–Discharge Rate on Lithium Reaction Distributions for a Commercial Lithium Coin Cell Visualized by Compton Scattering Imaging
Next Article in Special Issue
Anomalous Transport Behavior in Quantum Magnets
Previous Article in Journal
The Contribution of Synchrotron Light for the Characterization of Atmospheric Mineral Dust in Deep Ice Cores: Preliminary Results from the Talos Dome Ice Core (East Antarctica)
Previous Article in Special Issue
On the Evaluation of the Spin Galvanic Effect in Lattice Models with Rashba Spin-Orbit Coupling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Crossover Induced Electron Pairing and Superconductivity by Kinetic Renormalization in Correlated Electron Systems

by
Takashi Yanagisawa
1,*,
Mitake Miyazaki
2 and
Kunihiko Yamaji
1
1
National Institute of Advanced Industrial Science and Technology 1-1-1 Umezono, Tsukuba, Ibaraki 305-8568, Japan
2
Hakodate Institute of Technology, 14-1 Tokura, Hakodate, Hokkaido 042-8501, Japan
*
Author to whom correspondence should be addressed.
Condens. Matter 2018, 3(3), 26; https://doi.org/10.3390/condmat3030026
Submission received: 30 June 2018 / Revised: 22 August 2018 / Accepted: 5 September 2018 / Published: 6 September 2018
(This article belongs to the Special Issue Selected Papers from Quantum Complex Matter 2018)

Abstract

:
We investigate the ground state of strongly correlated electron systems based on an optimization variational Monte Carlo method to clarify the mechanism of high-temperature superconductivity. The wave function is optimized by introducing variational parameters in an exponential-type wave function beyond the Gutzwiller function. The many-body effect plays an important role as an origin of superconductivity in a correlated electron system. There is a crossover between weakly correlated region and strongly correlated region, where two regions are characterized by the strength of the on-site Coulomb interaction U. We insist that high-temperature superconductivity occurs in the strongly correlated region.

1. Introduction

The mechanism of high-temperature superconductivity has been studied vigorously for more than 30 years since the discovery of cuprate high-temperature superconductors [1]. High-temperature cuprates are typical strongly correlated systems since the parent materials are Mott insulators when no carriers are doped. It is important to understand electronic properties of strongly correlated electron systems.
The CuO 2 plane is commonly contained in high-temperature cuprates, where the CuO 2 plane consists of oxygen atoms and copper atoms. The electronic model for this plane is called the d-p model (or three-band Hubbard model) [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18]. It appears very hard to elucidate the phase diagram of the d-p model because of strong correlation between electrons. We often examine simplified models such as the two-dimensional single-band Hubbard model [19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38] or ladder model [39,40,41,42,43] as an attempt to make clear the phase diagram of correlated electron systems.
The Hubbard model is one of the fundamental models in the study of condensed matter physics. The Hubbard model is employed to understand the metal-insulator transition [44] and magnetic properties of various compounds [45,46]. The Hubbard model and also the d-p model have the potential to explain various phenomena. By employing the Hubbard model, we can understand the appearance of inhomogeneous states, reported for cuprates, such as stripes [47,48,49,50,51,52,53,54] and checkerboard-like density wave states [55,56,57,58].
It is important to clarify whether high-temperature superconductivity can be explained based on the Hubbard model (and the three-band d-p model). Previous works on the ladder Hubbard model supported the existence of superconducting phase [39,40,41,42,43,59]. Recent studies on the two-dimensional (2D) Hubbard model show positive results on superconductivity (SC) [38]. We show the order parameters of antiferromagnetic (AF) state and SC state as a function of the interaction parameter U in Figure 1. We think that the result strongly supports the existence of SC and shows a possibility of high-temperature superconductivity in the strongly correlated region.
A variational Monte Carlo method is a useful method to investigate electronic properties of strongly correlated electron systems by calculating the expectation values numerically [26,27,28,29,30,31]. A variational wave function can be improved by introducing new variational parameters to control the electron correlation. We have proposed wave functions that are optimized by multiplying an initial function by exp ( S ) -type operators [38,60,61], where S is a suitable correlation operator. The Gutzwiller function is also written in this form. An optimization process is performed in a systematic way by multiplying by the exponential-type operators repeatedly [60]. The ground-state energy is lowered considerably by using this type of wave functions [38]. This paper can be viewed as an extension of the paper published in proceeding of conference on superconductivity (ISS2017) [62].

2. Model Hamiltonians

The Hubbard model is written as
H = i j σ t i j c i σ c j σ + U i n i n i ,
where t i j indicates the transfer integral and U is the strength of the on-site Coulomb interaction. We set t i j = t when i and j are nearest-neighbor pairs i j and t i j = t when i and j are next-nearest neighbor pairs. We consider this model in two dimensions, and N and N e denote the number of lattice sites and the number of electrons, respectively.
We also consider the three-band model that explicitly contains oxygen p and copper d electrons. The Hamiltonian is
H d p = ϵ d i σ d i σ d i σ + ϵ p i σ ( p i + x ^ / 2 σ p i + x ^ / 2 σ + p i + y ^ / 2 σ p i + y ^ / 2 σ ) + t d p i σ [ d i σ ( p i + x ^ / 2 σ + p i + y ^ / 2 σ p i x ^ / 2 σ p i y ^ / 2 σ ) + h . c . ] + t p p i σ [ p i + y ^ / 2 σ p i + x ^ / 2 σ p i + y ^ / 2 σ p i x ^ / 2 σ p i y ^ / 2 σ p i + x ^ / 2 σ + p i y ^ / 2 σ p i x ^ / 2 σ + h . c . ] + t d i j σ ϵ i j ( d i σ d j σ + h . c . ) + U d i d i d i d i d i + U p i ( n i + x ^ / 2 p n i + x ^ / 2 p + n i + y ^ / 2 p n i + y ^ / 2 p ) .
We use the hole picture in this paper. d i σ and d i σ represent the operators for the d hole. p i ± x ^ / 2 σ and p i ± x ^ / 2 σ denote the operators for the p holes at the site R i ± x ^ / 2 , and in a similar way p i ± y ^ / 2 σ and p i ± y ^ / 2 σ are defined. n i + x ^ / 2 σ p and n i + y ^ / 2 σ p denote the number operators of p holes at R i + x ^ / 2 and R i + y ^ / 2 , respectively. t d p is the transfer integral between adjacent Cu and O orbitals and t p p is that between nearest p orbitals. t d indicates that between d orbitals where i j denotes a next nearest-neighbor pair of copper sites. cuprate superconductors such as Bi 2 Sr 2 CaCu 2 O 8 + δ [63] and Tl 2 ba 2 CuO 6 + δ [64]. ϵ i j takes the values ± 1 according to the sign of the transfer integral between next nearest-neighbor d orbitals. U d is the strength of the on-site Coulomb repulsion between d holes and U p is that between p holes. We can reproduce the Fermi surface by using parameters t d p , t p p and t d . The values of band parameters have been estimated [65,66,67,68,69]; for example, U d = 10.5 , U p = 4.0 and U d p = 1.2 in eV [66] where U d p is the nearest-neighbor Coulomb interaction between holes on adjacent Cu and O orbitals. We neglect U d p because U d p is small compared to U d . We use the notation Δ d p = ϵ p ϵ d . The number of sites is denoted as N, and the total number of atoms is N a = 3 N . The energy unit is given by t d p .

3. Optimization Variational Monte Carlo Method

In a variational Monte Carlo method, we employ a wave function that is suitable for the system which we consider and evaluate the expectation values by using a Monte Carlo procedure. To take into account electron correlation between electrons, we start from the Gutzwiller wave function given by
ψ G = P G ψ 0 ,
where P G is the Gutzwiller operator P G = j ( 1 ( 1 g ) n j n j ) where g is the variational parameter in the range of 0 g 1 . ψ 0 indicates a trial one-particle state.
Because the Gutzwiller function is very simple and is not enough to take account of electron correlation, we should improve the wave function. There are several methods to optimize the wave function. One method is to multiply the Gutzwiller function by an exponential-type operator. The wave function is written as [38,60,70,71,72,73,74]
ψ λ = exp ( λ K ) ψ G ,
where K is the kinetic part of the Hamiltonian and λ is a real variational operator [31,60,71]. The expectation values are calculated by using the auxiliary field method [60,75]. The other method to improve the Gutzwiller function is to introduce a Jastrow-type operator [33]. We can control nearest-neighbor correlation by multiplying by the operator
P J d h = j 1 ( 1 η ) τ [ d j ( 1 e j + τ ) + e j ( 1 d j + τ ) ] ,
where d j is the operator for the doubly-occupied site given as d j = n j n j and e j is that for the empty site given by e j = ( 1 n j ) ( 1 n j ) . η is the variational parameter in the range 0 η 1 . With this operator we can include the doublon-holon correlation:
ψ η = P J d h ψ G .
It is possible to generalize the Jastrow operator to consider long-range electron correlation by introducing new variational parameters [76].
In this paper we use the wave function of exponential type in Equation (4). We call this type of wave function the off-diagonal wave function since the off-diagonal correlation in the site representation is taken into account in this wave function. We believe that it is more important to consider off-diagonal electron correlation than diagonal electron correlation. In fact, the energy is further lowered when we employ the off-diagonal wave function [38].

4. Antiferromagnetic Crossover

The AF one-particle state ψ A F is given by the eigenfunction of the AF trial Hamiltonian given by
H A F = i j σ t i j c i σ c j σ Δ A F i σ ( 1 ) x i + y i σ n i σ ,
where Δ A F is the AF order parameter and ( x i , y i ) represents the coordinates of the site i. With ψ A F , the wave function is given as
ψ λ , A F = exp ( λ K ) P G ψ A F .
In general, the AF state is very stable in the Hubbard model near half-filling. Thus, it is important to control AF magnetic order so that the superconducting state is realized. The 2D Hubbard model and the three-band d-p model have similarity with respect to magnetic order. We found that the AF state is more stable in the d-p model than in the single-band Hubbard model.
The stability of the AF state depends mainly on the electron density n e , the interaction strength U, and the transfer integral t and long-range transfers in the single-band Hubbard model. The AF correlation is induced as U increases from zero in weakly correlated region and is maximized when U is of the order of the bandwidth, say at U = U c , when carriers are doped. When U becomes larger than U c , the AF correlation turns to decrease. In the region where U is extremely large, the AF correlation is suppressed to a small value by large fluctuation. This is shown in Figure 1. Thus, there is a crossover between weakly correlated region and strongly correlated region.
The transfer integral t also shows non-trivial effect on the stability of AF magnetic order. As t increases, the AF correlation increases. (We adopt that t is negative in this paper.) We show Δ A F as a function of 1 n e in Figure 2 for t = 0 and Figure 3 for t = 0.2 . When t is negative, the AF region expands up to about 20 percent doping where 1 n e 0.2 .
In the three-band d-p model, the AF correlation is stronger than that in the single-band Hubbard model. We introduced the transfer integral t d to control the strength of antiferromagnetism. We found that the AF correlation is reduced when t d increases, which is in contrast to the role of t for the Hubbard model. The Figure 4 shows the phase diagram of the ground state in the half-filled case in the t p p t d plane where AFI and PI indicate antiferromagnetic insulator state and paramagnetic insulator state, respectively. A phase transition occurs from the antiferromagnetic insulator to the paramagnetic insulator as parameters t p p and t d increase. A copper oxide which ia an insulator without hole doping may be in the AFI region of thia figure. For such copper oxides, | t d | may not be very large and we expect that the value of t d may be in the range | t d | 0.2 t d . As the density of holes increases, the role of t d and t p p will become important.

5. Correlated Superconductivity

The superconducting state is represented by the BCS wave function
ψ B C S = k ( u k + v k c k c k ) | 0 ,
with coefficients u k and v k that appear in the ratio u k / v k = Δ k / ( ξ k + ξ k 2 + Δ k 2 ) , where Δ k is the gap function with k dependence and ξ k = ϵ k μ is the dispersion relation of conduction electrons. We assume the d-wave symmetry for Δ k : Δ k = Δ S C ( cos k x cos k y ) . The Gutzwiller BCS state is formulated as
ψ G B C S = P N e P G ψ B C S ,
where P N e indicates the operator to extract the state with N e electrons. In this wave function the electron number is fixed and thus the chemical potential in ξ k is regarded as a variational parameter. In the formulation of ψ λ , we use the BCS wave function without fixing the total electron number, namely, without the operator P N e . The chemical potential μ in ξ k is not a variational parameter and is used to adjust the total electron number. The wave function is given as
ψ λ = e λ K P G ψ B C S .
We perform the electron-hole transformation for down-spin electrons:
d k = c k , d k = c k ,
and not for up-spin electrons: c k = c k . The electron pair operator c k c k denotes the hybridization operator c k d k in this formulation.
We show the phase diagram in Figure 5 where the condensation energies for SC and AF states are shown as a function of the hole density 1 n e for U / t = 18 and t = 0 . The condensation energy Δ E is defined as the energy lowering due to the inclusion of the order parameter:
Δ E = E ( Δ = 0 ) E ( Δ o p t ) ,
where Δ is the SC or AF order parameter and Δ o p t is its optimized value. We set t = 0 because t = 0 is the most optimal parameter value for superconductivity. In the optimum range for superconductivity, a pure d-wave superconducting state is realized.
As shown in Figure 5, there is a coexistence of superconductivity and antiferroamgnetism when the doping rate x < 0.09 . There is a possibility of coexistence in the underdoped region. This may be related to unusual metallic properties of cuprate superconductors in the underdiped region. We expect that the area of coexistence phase becomes smaller as the wave function is improved and optimized further by multiplying by exponential correlation operators.

6. Phase Separation

We discuss the phase separation in the 2D Hubbard model here. An existence of the phase separation has been pointed out recently. The ground-state energy E ( N e ) , where N e is the number of electrons, may exhibit a singular behavior near half-filling when the ground-state energy at half-filling is lowered extremely due to the AF order. The quantity δ 2 E ( N e ) [ E ( N e + δ N e ) 2 E ( N e ) + E ( N e δ N e ) ] / ( δ N e ) 2 , being proportional to the second derivative of the energy E ( N e ) with respect to the electron number, has a possibility to be negative for low hole doping. The phase separation occurs when δ 2 E ( N e ) is negative. In our optimized wave function, the phase separation is restricted to the range x 1 n e 0.06 for t = 0 and the phase separation disappears for negative t = 0.2 . We think that there is a possibility that the phase separation will disappear as the wave function is optimized further by multiplying by operators P G and exp ( λ K ) .

7. Summary

We investigated the ground-state properties of the two-dimensional Hubbard model by using the optimization variational Monte Carlo method. We used the exponential type wave function given in the form exp ( λ S ) with an appropriate operator S and a variational parameter λ . With our wave function, the ground-state energy is lowered greatly and the energy expectation value is lower than that obtained by any other wave function such as the Gutzwiller wave function and also several proposed wave functions with many variational parameters. The ground-state energy is lowered due to the kinetic-energy gain coming from exp ( λ K ) .
The antiferromagnetic state is very stable near half-filling (with no carriers) in the 2D Hubbard model the also the three-band d-p model. The AF correlation is suppressed as the doping rate of holes increases. As the strength of the on-site Coulomb interaction U changes, the crossover occurs between weakly correlated region and strongly correlated region. In the strongly correlated region, where U is larger than U c which is of the order of the bandwidth, the AF correlation is suppressed. A decrease of AF correlation indicates an increase in spin and charge fluctuation. This fluctuation is caused by an increase in kinetic energy and would induce electron pairing. We expect that this would cause high-temperature superconductivity.
Lastly we supplement the crossover. We expect that the crossover behavior is a universal phenomenon. The s-d model shows a crossover from weakly coupling to strongly coupling regions as the temperature decreases [77,78,79]. The logarithmic dependence of the resistivity appears due to an anomaly associated with the crossover [80,81,82,83,84,85]. The two-impurity Kondo problem also shows a crossover [86,87,88]. There may be a universal class that includes the Kondo effect, QCD [89], BCS-BEC crossover [90], Superconductivity, sine-Gordon model [91,92,93]. Fluctuations associated with the crossover may be represented by excitation modes that occur near a phase transition [94].

Author Contributions

Conceptualization, T.Y. and M.M.; Methodology, T.Y. and K.Y.; Validation, T.Y., M.M. and K.Y.; Writing-Original Draft Preparation, T.Y.

Funding

This work was supported by a Grant-in-Aid for Scientific Research from the Ministry of Education, Culture, Sports, Science and Technology of Japan (Grant No. 17K05559).

Acknowledgments

A part of computations was supported by the Supercomputer Center of the Institute for Solid State Physics, the University of Tokyo.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
VMCvariational Monte Carlo method
AFantiferromagnetic
SCsuperconductivity or superconducting
2Dtwo-dimensional
AFIantiferromagnetic insulator
PIparamagnetic insulator

References

  1. Bednorz, J.B.; Müller, K.A. Possible high Tc superconductivity in the Ba-La-Cu-O system. Z. Phys. 1986, B64, 189–193. [Google Scholar] [CrossRef]
  2. Emery, V.J. Theory of high-Tc superconductivity in oxides. Phys. Rev. Lett. 1987, 58, 2794. [Google Scholar] [CrossRef] [PubMed]
  3. Hirsch, J.E.; Loh, E.Y.; Scalapinom, D.J.; Tang, S. Pairing interaction in CuO clusters. Phys. Rev. B 1989, B39, 243. [Google Scholar] [CrossRef]
  4. Scalettar, R.T.; Scalapino, D.J.; Sugar, R.L.; White, S.R. Antiferromagnetic, charge-transfer, and pairing correlations in the three-band Hubbard model. Phys. Rev. B 1991, B44, 770. [Google Scholar] [CrossRef]
  5. Oguri, A.; Asahatam, T.; Maekawa, S. Gutzwiller wave function in the three-band Hubarf model: A variational Monte Carlo study. Phys. Rev. B 1994, B49, 6880. [Google Scholar] [CrossRef]
  6. Koikegami, S.; Yamada, K. Antiferromagnetic and superconducting correlations on the d-p model. J. Phys. Soc. Jpn. 2000, 69, 768–776. [Google Scholar] [CrossRef]
  7. Yanagisawa, T.; Koike, S.; Yamaji, K. Ground state of the three-band Hubbard model. Phys. Rev. B 2001, B64, 184509. [Google Scholar] [CrossRef]
  8. Koikegami, S.; Yanagisawa, T. Superconducting gap of the two-dimensional d-p model with small Ud. J. Phys. Soc. Jpn. 2001, 70, 3499–3502. [Google Scholar] [CrossRef]
  9. Yanagisawa, T.; Koike, S.; Yamaji, K. Lattice distortions, incommensurability, and stripes in the electronic model for high-Tc cuprates. Phys. Rev. B 2003, B67, 132408. [Google Scholar] [CrossRef]
  10. Koikegami, S.; Yanagisawa, T. Superconductivity in Sr2RuO4 mediated by Coulomb scattering. Phys. Rev. B 2003, B67, 134517. [Google Scholar] [CrossRef]
  11. Koikegami, S.; Yanagisawa, T. Superconductivity in multilayer perovskite. J. Phys. Soc. Jpn. 2006, 75, 034715. [Google Scholar] [CrossRef]
  12. Yanagisawa, T.; Miyazaki, M.; Yamaji, K. Incommensurate antiferromagnetism coexisting with superconductivity in two-dimensional d-p model. J. Phys. Soc. 2009, 78, 031706. [Google Scholar] [CrossRef]
  13. Weber, C.; Lauchi, A.; Mila, F.; Giamarchi, T. Orbital currents in extended Hubbard model of High-Tc cuprate superconductors. Phys. Rev. Lett. 2009, 102, 017005. [Google Scholar] [CrossRef] [PubMed]
  14. Lau, B.; Berciu, M.; Sawatzky, G.A. High spin polaron in lightly doped CuO2 planes. Phys. Rev. Lett. 2011, 106, 036401. [Google Scholar] [CrossRef] [PubMed]
  15. Weber, C.; Giamarchi, T.; Varma, C.M. Phase diagram of a three-orbital model for high-Tc cuprate superconductors. Phys. Rev. Lett. 2014, 112, 117001. [Google Scholar] [CrossRef] [PubMed]
  16. Avella, A.; Mancini, F.; Paolo, F.; Plekhanov, E. Emery vs Hubbard model for cuprate superconductors: A composite operator method study. Eur. Phys. J. 2013, B86, 265. [Google Scholar] [CrossRef]
  17. Ebrahimnejad, H.; Sawatzky, G.A.; Berciu, M. Differences between the insulating limit quasiparticles of one-band and three-band cuprate models. J. Phys. Cond. Matter 2016, 28, 105603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Tamura, S.; Yokoyama, H. Variational study of magnetic ordered state in d-p model. Phys. Procedia 2016, 81, 5. [Google Scholar] [CrossRef]
  19. Hubbard, J. Electron correlations in narrow energy bands. Proc. Roy. Soc. Lond. 1963, 276, 238–257. [Google Scholar] [CrossRef]
  20. Hubbard, J. Electron correlations in narrow energy bands III. Proc. Roy. Soc. Lond. 1964, 281, 401–419. [Google Scholar] [CrossRef]
  21. Gutzwiller, M.C. Effect of correlation on the ferromagnetism of transition metals. Phys. Rev. Lett. 1963, 10, 159. [Google Scholar] [CrossRef]
  22. Zhang, S.; Carlson, J.; Gubernatis, J.E. Constrained path Monte Carlo method for fermion ground states. Phys. Rev. B 1997, B55, 7464. [Google Scholar] [CrossRef]
  23. Zhang, S.; Carlson, J.; Gubernatis, J.E. Pairing correlation in the two-dimensional Hubbard model. Phys. Rev. Lett. 1997, 78, 4486. [Google Scholar] [CrossRef]
  24. Yanagisawa, T.; Shimoi, Y. Exact results in strongly correlated electrons. Int. J. Mod. Phys. 1996, B10, 3383. [Google Scholar] [CrossRef]
  25. Yanagisawa, T.; Shimoi, Y. Ground state of the Kondo-Hubbard model at half-filling. Phys. Rev. Lett. 1995, 74, 4939. [Google Scholar] [CrossRef] [PubMed]
  26. Nakanishi, T.; Yamaji, K.; Yanagisawa, T. Variational Monte Carlo indications of d-wave superconductivity in the two-dimensional Hubbard model. J. Phys. Soc. Jpn. 1997, 66, 294–297. [Google Scholar] [CrossRef]
  27. Yamaji, K.; Yanagisawa, T.; Nakanishi, T.; Koike, S. Variational Monte Carlo study on the superconductivity in the two-dimensional Hubbard model. Physica 1998, C304, 225–238. [Google Scholar] [CrossRef]
  28. Yamaji, K.; Yanagisawa, T.; Koike, S. Bulk limit of superconducting condensation energy in 2D Hubbard model. Physica 2000, 284, 415–416. [Google Scholar] [CrossRef]
  29. Yamaji, K.; Yanagisawa, T.; Miyazaki, M.; Kadono, R. Superconducting condensation energy of the two-dimensional Hubbard model in the large-negative-t’ region. J. Phys. Soc. Jpn. 2011, 80, 083702. [Google Scholar] [CrossRef]
  30. Hardy, T.M.; Hague, P.; Samson, J.H.; Alexandrov, A.S. Superconductivity in a Hubbard-Fröhlich model in cuprates. Phys. Rev. B 2009, B79, 212501. [Google Scholar] [CrossRef]
  31. Yanagisawa, T.; Miyazaki, M.; Yamaji, K. Correlated-electron systems and high-temperature superconductivity. J. Mod. Phys. 2013, 4, 33. [Google Scholar] [CrossRef]
  32. Bulut, N. dx2y2 superconductivity and the Hubbard model. Adv. Phys. 2002, 51, 1587–1667. [Google Scholar] [CrossRef]
  33. Yokoyama, H.; Tanaka, Y.; Ogata, M.; Tsuchiura, H. Crossover of superconducting properties and kinetic-energy gain in two-dimensional Hubbard model. J. Phys. Soc. Jpn. 2004, 73, 1119–1122. [Google Scholar] [CrossRef]
  34. Aimi, T.; Imada, M. Does simple two-dimensional Hubbard model account for high-Tc superconductivity in copper oxides? J. Phys. Soc. Jpn. 2007, 76, 113708. [Google Scholar] [CrossRef]
  35. Miyazaki, M.; Yanagisawa, T.; Yamaji, K. Diagonal stripe states in th elight-doping region in the two-dimensional Hubbard model. J. Phys. Soc. Jpn. 2004, 73, 1643–1646. [Google Scholar] [CrossRef]
  36. Yanagisawa, T. Phase diagram of the t-U2 Hamiltonian of the weak coupling Hubbard model. N. J. Phys. 2008, 10, 023014. [Google Scholar] [CrossRef]
  37. Yanagisawa, T. Enhanced pair correlation functions in the two-dimensional Hubbard model. N. J. Phys. 2013, 15, 033012. [Google Scholar] [CrossRef] [Green Version]
  38. Yanagisawa, T. Crossover from wealy to strongly correlated regions in the two-dimensional Hubbard model—Off-diagonal Monte Carlo studies of Hubbard model II. J. Phys. Soc. Jpn. 2016, 85, 114707. [Google Scholar] [CrossRef]
  39. Noack, R.M.; White, S.R.; Scalapino, D.J. The doped two-chain Hubbard model. Europhys. Lett. 1995, 30, 163. [Google Scholar] [CrossRef]
  40. Noack, R.M.; Bulut, N.; Scalapino, D.J.; Zacher, M.J. Enhanced dx2y2 pairing correlations in the two-leg Hubbard ladder. Phys. Rev. B 1997, B56, 7162. [Google Scholar] [CrossRef]
  41. Yamaji, K.; Shimoi, Y.; Yanagisawa, T. Superconductivity indications of the two-chain Hubbard model due to the two-band effect. Physica 1994, C235, 2221–2222. [Google Scholar] [CrossRef]
  42. Yanagisawa, T.; Shimoi, Y.; Yamaji, K. Superconducting phase of a two-chain Hubbard model. Phys. Rev. B 1995, B52, R3860. [Google Scholar] [CrossRef]
  43. Nakano, T.; Kuroki, K.; Onari, S. Superconductivity due to spin fluctuations originating from multiple Fermi surfaces in the double chain superconductor Pr2Ba4Cu7O15−δ. Phys. Rev. B 2007, B76, 014515. [Google Scholar] [CrossRef]
  44. Mott, N.F. Metal-Insulator Transitions; Taylor and Francis Ltd.: London, UK, 1974. [Google Scholar]
  45. Moriya, T. Spin Fluctuations in Itinerant Electron Magnetism; Springer-Verlag: Berlin, Germany, 1985. [Google Scholar]
  46. Yosida, K. Theory of Magnetism; Springer-Verlag: Berlin, Germany, 1996. [Google Scholar]
  47. Tranquada, J.M.; Axe, J.D.; Ichikawa, N.; Nakamura, Y.; Uchida, S.; Nachumi, B. Neutron-scattering study of stripe-phase order of holes and spins in La1.48Nd0.4Sr0.12CuO4. Phys. Rev. B 1996, B54, 7489. [Google Scholar] [CrossRef]
  48. Suzuki, T.; Goto, T.; Chiba, K.; Shinoda, T.; Fukase, T.; Kimura, H.; Yamada, K.; Ohashi, M.; Yamaguchi, Y. Observation of modulated magnetic long-range order in La1.88Sr0.12CuO4. Phys. Rev. B 1998, B57, R3229. [Google Scholar] [CrossRef]
  49. Yamada, K.; Lee, C.H.; Kurahashi, K.; Wada, J.; Wakimoto, S.; Ueki, S.; Kimura, H.; Endoh, Y.; Hosoya, S.; Shirane, G.; et al. Doping dependence of the spatially modulated dynamical spin correlations and the superconducting-transition temperature in La2−xSrxCuO4. Phys. Rev. B 1998, B57, 6165. [Google Scholar] [CrossRef]
  50. Arai, M.; Nishijima, T.; Endoh, Y.; Egami, T.; Tajima, S.; Tomimoto, K.; Shiohara, Y.; Takahashi, M.; Garrett, A.; Bennington, S.M. Incommensurate spin dynamics of underdoped superconductor YBa2Cu3Y6.7. Phys. Rev. Lett. 1999, 83, 608. [Google Scholar] [CrossRef]
  51. Mook, H.A.; Dai, P.; Dogan, F.; Hunt, R.D. One-dimensional nature of the magnetic fluctuations in YBa2Cu3O6.6. Nature 2000, 404, 729. [Google Scholar] [CrossRef] [PubMed]
  52. Wakimoto, S.; Birgeneau, R.J.; Kastner, M.A.; Lee, Y.S.; Erwin, R.; Gehring, P.M.; Lee, S.H.; Fujita, M.; Yamada, K.; Endoh, Y.; et al. Direct observation of a one-dimensional static spin modulation in insulating La1.95Sr0.05CuO4. Phys. Rev. B 2000, B61, 3699. [Google Scholar] [CrossRef]
  53. Bianconi, A.; Saini, N.L.; Lanzara, A.; Missori, M.; Rossetti, T.; Oyanagi, H.; Yamaguchi, H.; Oka, K.; Ito, T. Determination of the local lattice distortions in the CuO2 plane of La1.85Sr0.15CuO4. Phys. Rev. Lett. 1996, 76, 3412. [Google Scholar] [CrossRef] [PubMed]
  54. Bianconi, A. Quantum materials: shape resonances in superstripes. Nat. Phys. 2013, 9, 536. [Google Scholar] [CrossRef]
  55. Hoffman, J.E.; McElroy, K.; Lee, D.H.; Lang, K.M.; Eisaki, H.; Uchida, S.; Davis, J.C. Imaging quasiparticle interference in Bi2Sr2CaCu2O8+δ. Science 2002, 295, 466. [Google Scholar] [CrossRef] [PubMed]
  56. Wise, W.D.; Boyer, M.C.; Chatterjee, K.; Kondo, T.; Takeuchi, T.; Ikuta, H.; Wang, Y.; Hudson, E.W. Charge-density-wave origin of cuprate checkerboard visualized by scanning tunnelling microscopy. Nat. Phys. 2008, 4, 696. [Google Scholar] [CrossRef]
  57. Hanaguri, T.; Lupien, C.; Kohsaka, Y.; Lee, D.H.; Azuma, M.; Takano, M.; Takagi, H.; Davis, J.C. A checkerboard electronic crystal state in lightly hole-doped Ca2−xNaxCuO2Cl2. Nature 2004, 430, 1001. [Google Scholar] [CrossRef] [PubMed]
  58. Miyazaki, M.; Yanagisawa, T.; Yamaji, K. Checkerboard states in the two-dimensional Hubbard model with the Bi2212-type band. J. Phys. Soc. Jpn. 2009, 78, 043706. [Google Scholar] [CrossRef]
  59. Koike, S.; Yamaji, K.; Yanagisawa, T. Effect of the medium-range transfer energies to the superconductivity in the two-chain Hubbard model. J. Phys. Soc. Jpn. 1999, 68, 1657–1663. [Google Scholar] [CrossRef]
  60. Yanagisawa, T.; Koike, S.; Yamaji, K. Off-diagonal wave function Monte Carlo Studies of Hubbard model I. J. Phys. Soc. Jpn. 1998, 67, 3867–3874. [Google Scholar] [CrossRef]
  61. Yanagisawa, T.; Miyazaki, M. Mott transition in cuprate high-temperature superconductors. Eur. Phys. Lett. 2014, 107, 27004. [Google Scholar] [CrossRef] [Green Version]
  62. Yanagisawa, T.; Miyazaki, M.; Yamaji, K. Optimized wave funtion by kinetic renormalization effect in strongly correlated region of the three-band Hubbard model. J. Phys. Conf. Ser. 2018, 1054, 012017. [Google Scholar] [CrossRef]
  63. McElroy, K.; Simmonds, R.W.; Hoffman, J.E.; Lee, D.H.; Orenstein, J.; Eisaki, H.; Uchida, S.; Davis, J.C. Relating atomic-scale electronic phenomena to wave-like quasiparticle states in superconducting Bi2Sr2CaCu2O8+δ. Nature 2003, 422, 592. [Google Scholar] [CrossRef] [PubMed]
  64. Hussey, N.E.; Abdel-Jawad, M.; Carrington, A.; Mackenzie, A.P.; Balicas, L. A coherent three-dimensional Fermi surface in a high-transition-temperature superconductor. Nature 2003, 425, 814. [Google Scholar] [CrossRef] [PubMed]
  65. Weber, C.; Haule, K.; Kotliar, G. Critical weights and waterfalls in doped charge-transfer insulators. Phys. Rev. B 2008, B78, 134519. [Google Scholar] [CrossRef]
  66. Hybertsen, M.S.; Schlüter, M.; Christensen, N.E. Calculation of Coulomb-interaction parameter for La2CuO4 using a constrained-density-functional approach. Phys. Rev. B 1989, B39, 9028. [Google Scholar] [CrossRef]
  67. Eskes, H.; Sawatzky, G.A.; Feiner, L.F. Effective transfer for singlets formed by hole doping in the high-Tc superconductors. Physica 1989, C160, 424–430. [Google Scholar] [CrossRef]
  68. McMahan, A.K.; Annett, J.F.; Martin, R.M. Cuprate parameters from numerical Wannier functions. Phys. Rev. B 1990, B42, 6268. [Google Scholar] [CrossRef]
  69. Eskes, H.; Sawatzky, G. Single-, triple-, or multiplel-band Hubbard models. Phys. Rev. B 1991, B43, 119. [Google Scholar] [CrossRef]
  70. Otsuka, H. Variational Monte Carlo studies of the Hubbard model in one- and two-dimensions. J. Phys. Soc. Jpn. 1992, 61, 1645–1656. [Google Scholar] [CrossRef]
  71. Yanagisawa, T.; Koike, S.; Yamaji, K. d-wave state with multiplicative correlation factors for the Hubbard model. J. Phys. Soc. Jpn. 1999, 68, 3608–3614. [Google Scholar] [CrossRef]
  72. Eichenberger, D.; Baeriswyl, D. Superconductivity and antiferromagnetism in the-dimensional Hubbard model: A variational study. Phys. Rev. B 2007, B76, 180504. [Google Scholar] [CrossRef]
  73. Baeriswyl, D.; Eichenberger, D.; Menteshashvii, M. Variational ground states of the two-dimensional Hubbard model. N. J. Phys. 2009, 11, 075010. [Google Scholar] [CrossRef] [Green Version]
  74. Baeriswyl, D. Superconductivity in the repulsive Hubbards model. J. Supercond. Novel Magn. 2011, 24, 1157–1159. [Google Scholar] [CrossRef]
  75. Yanagisawa, T. Quantum Monte Carlo diagonalization for many-fermion systems. Phys. Rev. B 2007, B75, 224503. [Google Scholar] [CrossRef]
  76. Misawa, T.; Imada, M. Origin of high-Tc superconductivity in doped Hubbard models and their extensions: Roles of uniform charge fluctuations. Phys. Rev. B 2014, B90, 115137. [Google Scholar] [CrossRef]
  77. Kondo, J. The Physics of Dilute Magnetic Alloys; Cambridge University Press: Cambridge, UK, 2012. [Google Scholar]
  78. Yuval, G.; Anderson, P.W. Exact results for the Kondo problem: one-body theory and extension to finite temperature. Phys. Rev. B 1970, B1, 1522. [Google Scholar] [CrossRef]
  79. Wilson, K.G. The renormalization group: Critical phenomena and the Kondo problem. Rev. Mod. Phys. 1975, 47, 773. [Google Scholar] [CrossRef]
  80. Nagaoka, Y. Self-consistent treatment of Kondo’s effect in dilute alloys. Phys. Rev. B 1965, 138, A1112. [Google Scholar] [CrossRef]
  81. Zittartz, J.; Müller-Hartmann, E. Green’s function theory of the Kondo effect in dilute magnetic alloys. Z. Phys. 1968, 212, 380–407. [Google Scholar] [CrossRef]
  82. Anderson, P.W.; Yuval, G.; Hamann, D.R. Exact results in the Kondo problem II. Scaling theory, qualitatively correct solution, ans some new results on one-dimensional classical statistical models. Phys. Rev. B 1970, B1, 4464. [Google Scholar] [CrossRef]
  83. Yanagisawa, T. Kondo effect in the presence of spin-orbit coupling. J. Phys. Soc. Jpn. 2012, 81, 094713. [Google Scholar] [CrossRef]
  84. Yanagisawa, T. Kondo effect in Dirac systems. J. Phys. Soc. Jpn. 2015, 84, 074705. [Google Scholar] [CrossRef]
  85. Yanagisawa, T. Dirac fermions and Kondo effect. J. Phys. Conf. Ser. 2015, 603, 012014. [Google Scholar] [CrossRef] [Green Version]
  86. Jayaprakash, C.; Krishna-murthy, H.R.; Wilkins, J.W. Two-impurity Kondo problem. Phys. Rev. Lett. 1981, 47, 737. [Google Scholar] [CrossRef]
  87. Jones, B.A.; Varma, C.M. Critical point in the solution of the two-impurity Kondo problem. Phys. Rev. B 1989, B40, 324. [Google Scholar] [CrossRef]
  88. Yanagisawa, T. Ground state and staggered susceptibility of the two-impurity Kondo problem. J. Phys. Soc. Jpn. 1989, 60, 29–32. [Google Scholar] [CrossRef]
  89. Ellis, R.K.; Stirling, W.J.; Webber, B.R. QCD and Collider Physics; Cambridge University Press: Cambridge, UK, 1996. [Google Scholar]
  90. Nozieres, P.; Schmitt-Rink, S. Bose condensation in an attractive fermi gas: From weak to strong coupling superconductivity. J. Low Temp. Phys. 1985, 59, 195–211. [Google Scholar] [CrossRef]
  91. Rajaraman, R. Solitons and Instantons; North-Holland: Amsterdam, The Netherlands, 1989. [Google Scholar]
  92. Solyom, J. The Fermi gas model of one-dimensional conductors. Adv. Phys. 1979, 28, 201–303. [Google Scholar] [CrossRef]
  93. Yanagisawa, T. Chiral sine-Gordon model. Eur. Lett. 2016, 113, 41001. [Google Scholar] [CrossRef] [Green Version]
  94. Yanagisawa, T. Nambu-Goldstone bosons characterized by the order parameters in spontaneous symmetry breaking. J. Phys. Soc. Jpn. 2017, 86, 104711. [Google Scholar] [CrossRef]
Figure 1. AF and SC order parameters as a function of U / t on a 10 × 10 lattice with the periodic boundary condition in one direction and antiperiodic one in the other direction [38]. In Reference [38] Δ was shown as a function of U in the range 0 < U < 20 . We modified the figure to include the range 20 < U < 25 . Δ A F has a peak when U is of the order of the bandwidth U / t 10 . AF(G) in the figure shows the result obtainee for the simple Gutzwiller function.
Figure 1. AF and SC order parameters as a function of U / t on a 10 × 10 lattice with the periodic boundary condition in one direction and antiperiodic one in the other direction [38]. In Reference [38] Δ was shown as a function of U in the range 0 < U < 20 . We modified the figure to include the range 20 < U < 25 . Δ A F has a peak when U is of the order of the bandwidth U / t 10 . AF(G) in the figure shows the result obtainee for the simple Gutzwiller function.
Condensedmatter 03 00026 g001
Figure 2. Antiferromagnetic order parameters as a function of the hole density 1 n e on a 10 × 10 lattice for t = 0 . We put U / t = 12 , 14 and 18.
Figure 2. Antiferromagnetic order parameters as a function of the hole density 1 n e on a 10 × 10 lattice for t = 0 . We put U / t = 12 , 14 and 18.
Condensedmatter 03 00026 g002
Figure 3. Antiferromagnetic order parameters as a function of the hole density 1 n e on a 10 × 10 lattice for t = 0.2 t . We put U / t = 12 , 14 and 18.
Figure 3. Antiferromagnetic order parameters as a function of the hole density 1 n e on a 10 × 10 lattice for t = 0.2 t . We put U / t = 12 , 14 and 18.
Condensedmatter 03 00026 g003
Figure 4. Antiferromagnetic and paramagnetic insulator states in the plane of t p p and | t d | for the d-p model. We put U d = 8 , U p = 0 , ϵ p ϵ d = 1 and t d < 0 on a 6 × 6 lattice with 108 atoms in total. The energy unit is given by t d p .
Figure 4. Antiferromagnetic and paramagnetic insulator states in the plane of t p p and | t d | for the d-p model. We put U d = 8 , U p = 0 , ϵ p ϵ d = 1 and t d < 0 on a 6 × 6 lattice with 108 atoms in total. The energy unit is given by t d p .
Condensedmatter 03 00026 g004
Figure 5. The condensation energy per site as a function of the hole density x = 1 n e on a 10 × 10 lattice for the 2D Hubbard model. The SC and AF condensation energies are shown. We set t = 0 and U / t = 18 .
Figure 5. The condensation energy per site as a function of the hole density x = 1 n e on a 10 × 10 lattice for the 2D Hubbard model. The SC and AF condensation energies are shown. We set t = 0 and U / t = 18 .
Condensedmatter 03 00026 g005

Share and Cite

MDPI and ACS Style

Yanagisawa, T.; Miyazaki, M.; Yamaji, K. Crossover Induced Electron Pairing and Superconductivity by Kinetic Renormalization in Correlated Electron Systems. Condens. Matter 2018, 3, 26. https://doi.org/10.3390/condmat3030026

AMA Style

Yanagisawa T, Miyazaki M, Yamaji K. Crossover Induced Electron Pairing and Superconductivity by Kinetic Renormalization in Correlated Electron Systems. Condensed Matter. 2018; 3(3):26. https://doi.org/10.3390/condmat3030026

Chicago/Turabian Style

Yanagisawa, Takashi, Mitake Miyazaki, and Kunihiko Yamaji. 2018. "Crossover Induced Electron Pairing and Superconductivity by Kinetic Renormalization in Correlated Electron Systems" Condensed Matter 3, no. 3: 26. https://doi.org/10.3390/condmat3030026

Article Metrics

Back to TopTop