Next Article in Journal
Standard Behaviour of Bi2Sr2CaCu2O8+δ Overdoped
Previous Article in Journal
Z2 Topological Order and Topological Protection of Majorana Fermion Qubits
Previous Article in Special Issue
Potassium-Doped Para-Terphenyl: Structure, Electrical Transport Properties and Possible Signatures of a Superconducting Transition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On the Kinetic Energy Driven Superconductivity in the Two-Dimensional Hubbard Model

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.
Submission received: 16 October 2020 / Revised: 1 February 2021 / Accepted: 22 February 2021 / Published: 26 February 2021
(This article belongs to the Special Issue Quantum Complex Matter 2020)

Abstract

:
We investigate the role of kinetic energy for the stability of superconducting state in the two-dimensional Hubbard model on the basis of an optimization variational Monte Carlo method. The wave function is optimized by multiplying by correlation operators of site off-diagonal type. This wave function is written in an exponential-type form given as ψ λ = exp ( λ K ) ψ G for the Gutzwiller wave function ψ G and a kinetic operator K. The kinetic correlation operator exp ( λ K ) plays an important role in the emergence of superconductivity in large-U region of the two-dimensional Hubbard model, where U is the on-site Coulomb repulsive interaction. We show that the superconducting condensation energy mainly originates from the kinetic energy in the strongly correlated region. This may indicate a possibility of high-temperature superconductivity due to the kinetic energy effect in correlated electron systems.

1. Introduction

It is important and challenging to clarify the mechanism of high-temperature superconductivity in cuprates. It has been studied intensively for more than three decades [1]. In the study of cuprate high-temperature superconductivity, it is important to understand the ground state phase diagram. For this purpose, we should investigate electronic models with strong correlation.
It is certain that the CuO 2 plane plays an important role in the emergence of high-temperature superconductivity [2,3,4,5,6,7,8]. The CuO 2 plane contains oxygen atoms and copper atoms; thus, the CuO 2 plane can be modeled by the d-p model (or the three-band Hubbard model) [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25]. The single-band Hubbard model [26,27,28] is important since it can be regarded as a simplified model of the three-band d-p model and may contain important physics concerning high-temperature superconductivity [29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49]. The Hubbard model was first proposed by Hubbard to describe the metal-insulator transition [50] and has been one of the important fundamental models in condensed matter physics up to now. It may contain important physics concerning high-temperature cuprates, such as antiferromagnetic insulator, superconductivity, stripes [51,52,53,54,55,56,57,58], and inhomogeneous states [59,60,61,62].
We employ an optimization variational Monte Carlo method to investigate the ground state of the 2D Hubbard model. In a variational Monte Carlo method, we use variational wave functions with strong correlation between electrons [33,34,36,37,38,39]. A variational wave function is improved by introducing correlation operators. The wave function used in this paper is obtained by multiplying the Gutzwiller function by exp ( λ K ) operators, where K is the kinetic part of the Hamiltonian [47,48,63,64]. We can optimize the wave function further by multiplying by exponential-type operators again [63]. The ground-state energy is lowered greatly with this wave function [47].
The kinetic energy-driven mechanism of superconductivity has been examined in the study of cuprate superconductors for the Hubbard model [42,65,66,67] and the t-J model [68,69,70]. This is an interesting subject since there is a possibility that kinetic energy pairing occurs in high-temperature cuprates. In this paper, we discuss the role of kinetic energy based on the improved wave function for the Hubbard model. We evaluate the superconducting condensation energy as a sum of the kinetic and Coulomb energy contributions, and we show that the kinetic energy contribution dominates the condensation energy in the strongly correlated region of large U.

2. Hubbard Hamiltonian

We investigate the two-dimensional Hubbard model. The Hamiltonian of the Hubbard model is
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. The transfer integral is t i j = t when i and j are nearest-neighbor pairs i j . We put t = 0 in this paper, where t i j = t when i and j are next-nearest neighbor pairs. N and N e denote the number of lattice sites and the number of electrons, respectively. The energy unit is given by t. We define the non-interacting part as K:
K = i j σ t i j c i σ c j σ .
When two electrons with spin up and down are at the same site, the energy becomes higher by U. This simple interaction may cause many interesting phenomena, such as the metal-insulator transition, antiferromagnetic magnetism, and superconductivity.
The metal-insulator transition occurs at half filling when U is as large as the bandwidth. The effective Hamiltonian is given by the Heisenberg model when U is large in the half-filled case, which leads to the t-J model when holes are doped. Magnetic properties of materials may be described by the Hubbard model by introducing suitable magnetic orders. We discuss superconductivity in the strongly correlated region in this paper. The emergence of superconducting state in this region is closely related to the kinetic energy gain that increases as U increases.

3. Optimized Wave Function

In a variational Monte Carlo method, we calculate the expectation values of physical properties by using a Monte Carlo procedure. We start with the Gutzwiller wave function to take account of electron correlation. The Gutzwiller wave function is
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 one-particle state, such as the Fermi sea, the BCS state, and the antiferromagnetic state.
We improve the wave function by multiplying by the correlation operator given as [47,63,71,72,73,74,75]
ψ λ = exp ( λ K ) ψ G ,
where K is the kinetic part of the Hamiltonian, and λ is a real variational operator [39,63,72].
There are other methods to improve the Gutzwiller wave function by using Jastrow-type operators [41,76,77]. This operator is written as
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 that takes the value in the range of 0 η 1 . τ indicates a vector connecting nearest-neighbor sites. The Jastrow-type wave function is written as ψ J = P J d h ψ G . An important difference between ψ λ and ψ J is that P J d h is the site-diagonal operator, while ψ λ is the site-off diagonal operator. The one-particle state ψ 0 is written in the form
ψ 0 = j a j 0 φ j 0 ,
where { φ j 0 } is a set of basis functions of the one-particle state in the site representation on a lattice, with j being the label for the electron configuration. ψ λ and ψ J are given as
ψ λ = j a j 0 e λ K P G φ j 0 ,
ψ J = j a j 0 P J d h P G φ j 0 .
Since P G and P J d h are diagonal operators, ψ J is written as
ψ J = j a j J φ j 0 ,
where a j J = a j J ( g , η ) is determined by parameters g and η . ψ J is given by a wave function, where the coefficients { a j 0 } are modified in the one-particle state. Instead, ψ λ is not so simple because e λ K generates other basis states from φ j 0 , which means that off-diagonal elements φ i 0 K φ j 0 are effectively taken into account. We write ψ λ as
ψ λ = j a j λ φ j ,
where the set of basis states { φ j } may contain basis states which are not included in { φ j 0 } since some coefficients a 0 s may vanish accidentally.
The expectation values are calculated numerically by using the auxiliary field method following a Monte Carlo algorithm [63,78]. We show the ground-state energy per site as a function of U for ψ G and ψ λ in Figure 1. The energy is lowered due to the exponential factor e λ K . In the region of large U, the energy lowering mainly comes from the kinetic energy gain, as shown later.

4. Correlated Superconducting State

The correlated superconducting state is obtained by multiplying the BCS wave function by a correlation operator. The BCS wave function is
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 adopt the d-wave symmetry for Δ k , and we do not consider other symmetries in this paper since an SC state with other symmetry unlikely becomes stable in the simple single-band Hubbard model. The Gutzwiller BCS state is given by
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. Here, the total electron number is fixed, and the chemical potential in ξ k is regarded as a variational parameter.
The improved superconducting wave function is written as
ψ λ = e λ K P G ψ B C S .
In the formulation of ψ λ , we cannot fix the total electron number, and we should use a different Monte Carlo sampling procedure. 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 . In the real space, we have c i = c i and d i = c i .
The electron pair operator c k c k is transformed to the hybridization operator c k d k . Then, we can use the auxiliary field method after this transformation in a Monte Carlo simulation. This is performed by expressing the Gutzwiller operator in the form [72].
P G = i ( 1 ( 1 g ) c i c i ( 1 d i d i ) ) = i exp ( α c i c i + α c i c i d i d i ) = 1 2 N s i = ± 1 exp i ( 2 a s i α / 2 ) ( c i c i d i d i ) ,
where g = e α , cosh ( 2 a ) = e α / 2 , and s i is the auxiliary field that takes the value of ± 1 . By using this form, the expectation value is calculated as a sum of terms with respect to auxiliary fields, for which we apply the Monte Carlo procedure [63,72].

5. Kinetic Energy in the Superconducting State

In this section, we discuss the role of kinetic energy in the superconducting state in the two-dimensional Hubbard model. We consider the large-U region, where the kinetic energy of electrons would play an important role. Here, large-U means that U is much larger than the band width. The ground-state energy is determined by the balance of the kinetic energy and the Coulomb energy. We show the energy expectation values as a function of λ in Figure 2, where E g / N is the ground-state energy per site, E k i n / N is the expectation value of the non-interacting part of the Hamiltonian E k i n = ψ λ K ψ λ / ψ λ ψ λ , and E U denotes the Coulomb energy given by E U = U ψ λ i n i n i ψ λ / ψ λ ψ λ . E g has a minimum value for a finite value of λ .
The kinetic energy part gives a large contribution to E g when U is large. This is shown in Figure 3. When U > 10 t , E U for ψ λ almost agrees with that for ψ G . The difference of E k i n for ψ λ and ψ G increases for U > 10 t . The region that may be called the ’kinetic energy phase’ exists when 10 < U / t .
Let us consider the the difference of the kinetic energy defined as
Δ E k i n = E k i n ( ψ G ) E k i n ( ψ λ ) ,
where E k i n ( ψ G ) and E k i n ( ψ λ ) denote the kinetic energy for ψ G and ψ λ , respectively. Since ψ G = ψ λ = 0 with vanishing λ , we can write Δ E k i n = E k i n ( λ = 0 ) E k i n ( λ ) for the optimized value λ . We show Δ E k i n / N as a function of U in Figure 4, where the hole doping rate x is x = 0.12 . Δ E k i n begins to increase when U is of the order of the band width U 8 t . Δ E k i n has a broad peak when 15 t < U < 20 t . We define the SC condensation energy Δ E s c and the kinetic energy condensation energy Δ E k i n s c as
Δ E s c = E g ( Δ = 0 ) E g ( Δ = Δ opt ) ,
Δ E k i n s c = E k i n ( Δ = 0 ) E k i n ( Δ = Δ opt ) ,
where Δ = Δ s c is the superconducting order parameter, and Δ opt is its optimized value. We assumed the d-wave symmetry for Δ k : Δ k = Δ ( cos k x cos k y ) . Δ E k i n s c / N is also shown in Figure 4. The figure indicates that Δ E k i n s c increases for large U, showing a similar behavior to Δ E k i n . Δ E k i n s c can change the sign when U is small, which is consistent with the analysis for Bi 2 Sr 2 CaCu 2 O 8 + δ [79].
In Figure 5, we show Δ E k i n / N and Δ E k i n s c / N for x = 0.20 . Δ E k i n for x = 0.20 is smaller than that for x = 0.12 . The kinetic energy condensation energy Δ E k i n s c is also reduced for x = 0.20 compared to that for x = 0.12 . This result inevitably leads to the decrease of the SC condensation energy [48]. Hence, the kinetic energy effect becomes weak when the hole doping rate is large.
The Coulomb energy gain in the presence of the SC order parameter Δ E U s c is also evaluated, where
Δ E U s c = E U ( Δ = 0 ) E U ( Δ = Δ opt ) .
Δ E U s c is the Coulomb energy contribution to the SC condensation energy; namely, we have
Δ E s c = Δ E k i n s c + Δ E U s c .
Our result shows that
Δ E k i n s c > 0 , Δ E U s c < 0 ,
for the improved state ψ λ B C S . We show the SC condensation energy Δ E s c ( Δ s c ) and the Coulomb energy part Δ E U s c ( Δ s c ) as a function of Δ s c in Figure 6 where Δ E s c ( Δ ) = E g ( Δ = 0 ) E g ( Δ ) and Δ E U s c ( Δ ) = E U ( Δ = 0 ) E U ( Δ ) . The positive Δ E k i n s c > 0 indicates that the SC state ψ λ B C S becomes stable due to the kinetic energy effect.

6. Summary

We investigated the ground state of the two-dimensional Hubbard model by using the optimization variational Monte Carlo method with focus on the strongly correlated large-U region. The ground-state energy is greatly lowered due to the exp ( λ K ) correlation operator. The optimization variational Monte Carlo method is effective even in strongly correlated regions where U is much larger than the bandwidth.
In the large-U region, the kinetic energy term becomes dominant in the ground state, and the kinetic energy effect dominates the antiferromagnetic correlation. The kinetic energy difference Δ E k i n = E k i n ( ψ G ) E k i n ( ψ λ ) increases when U > 10 t and has a broad peak. The kinetic condensation energy Δ E k i n s c behaves like the kinetic energy difference Δ E k i n for U > 10 t . There is a correlation between Δ E k i n s c and Δ E k i n . The condensation energy Δ E s c mainly comes from the kinetic energy part Δ E k i n s c and there is a competition between the kinetic energy and the Coulomb energy. Hence there are two competitions in the Hubbard model; one is that between superconductivity and antiferromagnetism and the other is that between kinetic energy effect and Coulomb repulsive interaction. As a result of competitions, superconducting transition occurs. The result shows that superconductivity in the strongly correlated region is induced by kinetic energy effect. We expect that high-temperature superconductivity would be realized in the strongly correlated region of the two-dimensional Hubbard model.

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. All authors have read and agreed to the published version of the manuscript.

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. [Google Scholar] [CrossRef]
  2. 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]
  3. 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]
  4. Weber, C.; Haule, K.; Kotliar, G. Critical weights and waterfalls in doped charge-transfer insulators. Phys. Rev. 2008, B78, 134519. [Google Scholar] [CrossRef] [Green Version]
  5. Hybertsen, M.S.; Schlüter, M.; Christensen, N.E. Calculation of Coulomb-interaction parameter for La2CuO4 using a constrained-density-functional approach. Phys. Rev. 1989, B39, 9028. [Google Scholar] [CrossRef] [PubMed]
  6. 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. [Google Scholar] [CrossRef]
  7. McMahan, A.K.; Annett, J.F.; Martin, R.M. Cuprate parameters from numerical Wannier functions. Phys. Rev. 1990, B42, 6268. [Google Scholar] [CrossRef] [PubMed]
  8. Eskes, H.; Sawatzky, G. Single-, triple-, or multiple-band Hubbard models. Phys. Rev. 1991, B43, 119. [Google Scholar] [CrossRef] [PubMed]
  9. Emery, V.J. Theory of high-Tc superconductivity in oxides. Phys. Rev. Lett. 1987, 58, 2794. [Google Scholar] [CrossRef] [Green Version]
  10. Hirsch, J.E.; Loh, E.Y.; Scalapinom, D.J.; Tang, S. Pairing interaction in CuO clusters. Phys. Rev. 1989, B39, 243. [Google Scholar] [CrossRef]
  11. 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. 1991, B44, 770. [Google Scholar] [CrossRef]
  12. Oguri, A.; Asahatam, T.; Maekawa, S. Gutzwiller wave function in the three-band Hubarf model: A variational Monte Carlo study. Phys. Rev. 1994, B49, 6880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Koikegami, S.; Yamada, K. Antiferromagnetic and superconducting correlations on the d-p model. J. Phys. Soc. Jpn. 2000, 69, 768. [Google Scholar] [CrossRef]
  14. Yanagisawa, T.; Koike, S.; Yamaji, K. Ground state of the three-band Hubbard model. Phys. Rev. 2001, B64, 184509. [Google Scholar] [CrossRef] [Green Version]
  15. Koikegami, S.; Yanagisawa, T. Superconducting gap of the two-dimensional d-p model with small Ud. J. Phys. Soc. Jpn. 2001, 70, 3499. [Google Scholar] [CrossRef]
  16. Yanagisawa, T.; Koike, S.; Yamaji, K. Lattice distortions, incommensurability, and stripes in the electronic model for high-Tc cuprates. Phys. Rev. 2003, B67, 132408. [Google Scholar] [CrossRef] [Green Version]
  17. Koikegami, S.; Yanagisawa, T. Superconductivity in Sr2RuO4 mediated by Coulomb scattering. Phys. Rev. 2003, B67, 134517. [Google Scholar] [CrossRef] [Green Version]
  18. Koikegami, S.; Yanagisawa, T. Superconductivity in multilayer perovskite. J. Phys. Soc. Jpn. 2006, 75, 034715. [Google Scholar] [CrossRef] [Green Version]
  19. 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]
  20. 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] [Green Version]
  21. 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]
  22. 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] [Green Version]
  23. 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] [Green Version]
  24. 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] [Green Version]
  25. Tamura, S.; Yokoyama, H. Variational study of magnetic ordered state in d-p model. Phys. Procedia 2016, 81, 5. [Google Scholar] [CrossRef] [Green Version]
  26. Hubbard, J. Electron correlations in narrow energy bands. Proc. R. Soc. Lond. 1963, 276, 238. [Google Scholar]
  27. Hubbard, J. Electron correlations in narrow energy bands III. Proc. R. Soc. Lond. 1964, 281, 401. [Google Scholar]
  28. Gutzwiller, M.C. Effect of correlation on the ferromagnetism of transition metals. Phys. Rev. Lett. 1963, 10, 159. [Google Scholar] [CrossRef]
  29. Zhang, S.; Carlson, J.; Gubernatis, J.E. Constrained path Monte Carlo method for fermion ground states. Phys. Rev. 1997, B55, 7464. [Google Scholar] [CrossRef] [Green Version]
  30. Zhang, S.; Carlson, J.; Gubernatis, J.E. Pairing correlation in the two-dimensional Hubbard model. Phys. Rev. Lett. 1997, 78, 4486. [Google Scholar] [CrossRef] [Green Version]
  31. Yanagisawa, T.; Shimoi, Y. Exact results in strongly correlated electrons. Int. J. Mod. Phys. 1996, B10, 3383. [Google Scholar] [CrossRef]
  32. Yanagisawa, T.; Shimoi, Y.; Yamaji, K. Superconducting phase of a two-chain Hubbard model. Phys. Rev. 1995, B52, R3860. [Google Scholar] [CrossRef] [PubMed]
  33. 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. [Google Scholar] [CrossRef]
  34. 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. [Google Scholar] [CrossRef] [Green Version]
  35. 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. [Google Scholar] [CrossRef]
  36. Yamaji, K.; Yanagisawa, T.; Koike, S. Bulk limit of superconducting condensation energy in 2D Hubbard model. Physica 2000, B284, 415–416. [Google Scholar] [CrossRef]
  37. 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] [Green Version]
  38. Hardy, T.M.; Hague, P.; Samson, J.H.; Alexandrov, A.S. Superconductivity in a Hubbard-Fröhlich model in cuprates. Phys. Rev. 2009, B79, 212501. [Google Scholar] [CrossRef] [Green Version]
  39. Yanagisawa, T.; Miyazaki, M.; Yamaji, K. Correlated-electron systems and high-temperature superconductivity. J. Mod. Phys. 2013, 4, 33. [Google Scholar] [CrossRef] [Green Version]
  40. Bulut, N. dx2−y2 superconductivity and the Hubbard model. Adv. Phys. 2002, 51, 1587. [Google Scholar] [CrossRef] [Green Version]
  41. 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. [Google Scholar] [CrossRef] [Green Version]
  42. Yokoyama, H.; Ogata, M.; Tanaka, Y. Mott transitions and d-wave superconductivity in half-filled Hubbard model on square lattice with geometric frustration. J. Phys. Soc. Jpn. 2006, 75, 114706. [Google Scholar] [CrossRef] [Green Version]
  43. 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] [Green Version]
  44. Miyazaki, M.; Yanagisawa, T.; Yamaji, K. Diagonal stripe states in the light-doping region in the two-dimensional Hubbard model. J. Phys. Soc. Jpn. 2004, 73, 1643. [Google Scholar] [CrossRef]
  45. Yanagisawa, T. Phase diagram of the t-U2 Hamiltonian of the weak coupling Hubbard model. New J. Phys. 2008, 10, 023014. [Google Scholar] [CrossRef] [Green Version]
  46. Yanagisawa, T. Enhanced pair correlation functions in the two-dimensional Hubbard model. New J. Phys. 2013, 15, 033012. [Google Scholar] [CrossRef] [Green Version]
  47. 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] [Green Version]
  48. Yanagisawa, T. Antiferromagnetic, Superconductivity and phase diagram in the two-dimensional Hubbard model-Off-diagonal wave function Monte Carlo studies of Hubbard model. J. Phys. Soc. Jpn. 2019, 88, 054702. [Google Scholar] [CrossRef]
  49. Yanagisawa, T. Mechanism of high-temperature superconductivity in correlated-electron systems. Condens. Matter 2019, 4, 57. [Google Scholar] [CrossRef] [Green Version]
  50. Mott, N.F. Metal-Insulator Transitions; Taylor and Francis Ltd.: London, UK, 1974. [Google Scholar]
  51. 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. 1996, B54, 7489. [Google Scholar] [CrossRef] [PubMed]
  52. 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. 1998, B57, R3229. [Google Scholar] [CrossRef]
  53. 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. 1998, B57, 6165. [Google Scholar] [CrossRef]
  54. 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]
  55. 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] [Green Version]
  56. 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. 2000, B61, 3699. [Google Scholar] [CrossRef] [Green Version]
  57. 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] [Green Version]
  58. Bianconi, A. Quantum materials: Shape resonances in superstripes. Nat. Phys. 2013, 9, 536. [Google Scholar] [CrossRef]
  59. 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] [Green Version]
  60. 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]
  61. 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]
  62. 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] [Green Version]
  63. Yanagisawa, T.; Koike, S.; Yamaji, K. Off-diagonal wave function Monte Carlo Studies of Hubbard model I. J. Phys. Soc. Jpn. 1998, 67, 3867. [Google Scholar] [CrossRef] [Green Version]
  64. Yanagisawa, T.; Miyazaki, M. Mott transition in cuprate high-temperature superconductors. EPL 2014, 107, 27004. [Google Scholar] [CrossRef] [Green Version]
  65. Maier, T.A.; Jarrell, M.; Macridin, A.; Slezak, C. Kinetic energy driven pairing in cuprate superconductors. Phys. Rev. Lett. 2004, 92, 027005. [Google Scholar] [CrossRef] [Green Version]
  66. Gull, E.; Millis, A.J. Energetics of superconductivity in the two-dimensional Hubbard model. Phys. Rev. B 2012, 86, 241106. [Google Scholar] [CrossRef] [Green Version]
  67. Tocchio, L.F.; Becca, F.; Sorella, S. Hidden Mott transition and large-U superconductivity in the two-dimensional Hubbard model. Phys. Rev. B 2016, 94, 195126. [Google Scholar] [CrossRef] [Green Version]
  68. Feng, S. Kinetic energy driven superconductivity in doped cuprates. Phys. Rev. B 2003, 68, 184501. [Google Scholar] [CrossRef] [Green Version]
  69. Wrobel, P.; Eder, R.; Micnas, R. Kinetic energy driven superconductivity and the pseudogap phase in weakly doped antiferromagnets. J. Phys. Condens. Matter 2003, 15, 2755. [Google Scholar] [CrossRef] [Green Version]
  70. Guo, H.; Feng, S. Electronic structure of kinetic energy driven superconductors. Phys. Lett. A 2007, 361, 382. [Google Scholar] [CrossRef] [Green Version]
  71. Otsuka, H. Variational Monte Carlo studies of the Hubbard model in one- and two-dimensions. J. Phys. Soc. Jpn. 1992, 61, 1645. [Google Scholar] [CrossRef]
  72. Yanagisawa, T.; Koike, S.; Yamaji, K. d-wave state with multiplicative correlation factors for the Hubbard model. J. Phys. Soc. Jpn. 1999, 68, 3608. [Google Scholar] [CrossRef]
  73. Eichenberger, D.; Baeriswyl, D. Superconductivity and antiferromagnetism in the-dimensional Hubbard model: A variational study. Phys. Rev. 2007, B76, 180504. [Google Scholar] [CrossRef] [Green Version]
  74. Baeriswyl, D.; Eichenberger, D.; Menteshashvii, M. Variational ground states of the two-dimensional Hubbard model. New J. Phys. 2009, 11, 075010. [Google Scholar] [CrossRef]
  75. Baeriswyl, D. Superconductivity in the repulsive Hubbards model. J. Supercond. Novel Magn. 2011, 24, 1157. [Google Scholar] [CrossRef] [Green Version]
  76. Capello, M.; Becca, F.; Fabrizio, M.; Sorella, S.; Tosatti, E. Variational description of Mott insulators. Phys. Rev. Lett. 2005, 94, 026406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Misawa, T.; Imada, M. Origin of high-Tc superconductivity in doped Hubbard models and their extensions: Roles of uniform charge fluctuations. Phys. Rev. 2014, B90, 115137. [Google Scholar] [CrossRef] [Green Version]
  78. Yanagisawa, T. Quantum Monte Carlo diagonalization for many-fermion systems. Phys. Rev. 2007, B75, 224503. [Google Scholar] [CrossRef] [Green Version]
  79. Deutscher, G.; Santander-Syro, A.F.; Bontemps, N. Kinetic energy change with doping upon superfluid condensation in high-temperature superconductors. Phys. Rev. 2005, B72, 092504. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Ground-state energy as a function of U on a 10 × 10 lattice, where we set N e = 88 and t = 0 with the boundary condition in one direction, and the antiperiodic one in the other direction.
Figure 1. Ground-state energy as a function of U on a 10 × 10 lattice, where we set N e = 88 and t = 0 with the boundary condition in one direction, and the antiperiodic one in the other direction.
Condensedmatter 06 00012 g001
Figure 2. Ground-state energy E g / N , kinetic energy E k i n / N , and the expectation value of the Coulomb interaction E U / N as a function of λ on a 10 × 10 lattice with the periodic boundary condition in one direction, and the antiperiodic one in the other direction. We set N e = 88 , U = 18 t and t = 0 .
Figure 2. Ground-state energy E g / N , kinetic energy E k i n / N , and the expectation value of the Coulomb interaction E U / N as a function of λ on a 10 × 10 lattice with the periodic boundary condition in one direction, and the antiperiodic one in the other direction. We set N e = 88 , U = 18 t and t = 0 .
Condensedmatter 06 00012 g002
Figure 3. Kinetic energy E k i n / N and the Coulomb energy E U / N as a function of U on a 10 × 10 lattice. The boundary conditions are the same as in Figure 2. The electron number is N e = 88 , and we put t = 0 . Open circles indicate the results for the Gutzwiller function, and solid circles are those for the improved ψ λ .
Figure 3. Kinetic energy E k i n / N and the Coulomb energy E U / N as a function of U on a 10 × 10 lattice. The boundary conditions are the same as in Figure 2. The electron number is N e = 88 , and we put t = 0 . Open circles indicate the results for the Gutzwiller function, and solid circles are those for the improved ψ λ .
Condensedmatter 06 00012 g003
Figure 4. Kinetic-energy difference Δ E k i n / N and the kinetic energy difference in the SC state Δ E k i n s c / N ( × 25 ) as a function of U on a 10 × 10 lattice. Since Δ E k i n s c / N is small compared to Δ E k i n / N , we multiplied Δ E k i n s c / N by 25. The boundary conditions are the same as in Figure 1. The electron number is N e = 88 ( n = 0.88 ) and we put t = 0 .
Figure 4. Kinetic-energy difference Δ E k i n / N and the kinetic energy difference in the SC state Δ E k i n s c / N ( × 25 ) as a function of U on a 10 × 10 lattice. Since Δ E k i n s c / N is small compared to Δ E k i n / N , we multiplied Δ E k i n s c / N by 25. The boundary conditions are the same as in Figure 1. The electron number is N e = 88 ( n = 0.88 ) and we put t = 0 .
Condensedmatter 06 00012 g004
Figure 5. Kinetic-energy difference Δ E k i n / N and the kinetic energy difference in the SC state Δ E k i n s c / N ( × 25 ) as a function of U on a 10 × 10 lattice for N e = 80 ( n = 0.80 ) and t = 0 . The boundary conditions are the same as in Figure 1.
Figure 5. Kinetic-energy difference Δ E k i n / N and the kinetic energy difference in the SC state Δ E k i n s c / N ( × 25 ) as a function of U on a 10 × 10 lattice for N e = 80 ( n = 0.80 ) and t = 0 . The boundary conditions are the same as in Figure 1.
Condensedmatter 06 00012 g005
Figure 6. SC condensation energy Δ E s c ( Δ s c ) / N and its Coulomb energy part | Δ E U s c ( Δ s c ) / N | as a function of Δ s c on a 10 × 10 lattice for N e = 88 ( n = 0.88 ) and U = 18 . We set t = 0 . Δ E U s c is negative for this set of parameters. The boundary conditions are the same as in Figure 1.
Figure 6. SC condensation energy Δ E s c ( Δ s c ) / N and its Coulomb energy part | Δ E U s c ( Δ s c ) / N | as a function of Δ s c on a 10 × 10 lattice for N e = 88 ( n = 0.88 ) and U = 18 . We set t = 0 . Δ E U s c is negative for this set of parameters. The boundary conditions are the same as in Figure 1.
Condensedmatter 06 00012 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yanagisawa, T.; Yamaji, K.; Miyazaki, M. On the Kinetic Energy Driven Superconductivity in the Two-Dimensional Hubbard Model. Condens. Matter 2021, 6, 12. https://0-doi-org.brum.beds.ac.uk/10.3390/condmat6010012

AMA Style

Yanagisawa T, Yamaji K, Miyazaki M. On the Kinetic Energy Driven Superconductivity in the Two-Dimensional Hubbard Model. Condensed Matter. 2021; 6(1):12. https://0-doi-org.brum.beds.ac.uk/10.3390/condmat6010012

Chicago/Turabian Style

Yanagisawa, Takashi, Kunihiko Yamaji, and Mitake Miyazaki. 2021. "On the Kinetic Energy Driven Superconductivity in the Two-Dimensional Hubbard Model" Condensed Matter 6, no. 1: 12. https://0-doi-org.brum.beds.ac.uk/10.3390/condmat6010012

Article Metrics

Back to TopTop