This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Resonant multi-gap superconductivity at room temperature near a Lifshitz topological transition in sulfur hydrides

Maria Vittoria Mazziotti RICMASS Rome International Center for Materials Science, Superstripes Via dei Sabelli 119A, 00185 Roma, Italy Department of Mathematics and Physics, University Roma Tre, via della Vasca Navale 84, 00146 Roma, Italy    Roberto Raimondi Department of Mathematics and Physics, University Roma Tre, via della Vasca Navale 84, 00146 Roma, Italy    Antonio Valletta Italian National Research Council CNR, Institute for Microelectronics and Microsystems IMM, via del Fosso del Cavaliere, 100, 00133 Roma, Italy    Gaetano Campi Institute of Crystallography, CNR, via Salaria Km 29.300, I-00015 Monterotondo, Roma, Italy    Antonio Bianconi RICMASS Rome International Center for Materials Science, Superstripes Via dei Sabelli 119A, 00185 Roma, Italy Institute of Crystallography, CNR, via Salaria Km 29.300, I-00015 Monterotondo, Roma, Italy National Research Nuclear University MEPhI (Moscow Engineering Physics Institute), 115409 Moscow, Russia
Abstract

The maximum critical temperature for superconductivity in pressurized hydrides appears at the top of superconducting domes in TcT_{c} versus pressure curves at a particular pressure, which is not predicted by standard superconductivity theories. The a high-order anisotropic van Hove singularity near the Fermi level observed in band structure calculations of pressurized sulfur hydride, typical of a supermetal, has been associated with the array of metallic hydrogen wires modules forming a nanoscale heterostructure at atomic limit called superstripes phase. Here we propose that pressurized sulfur hydrides behave as a heterostructure made of a nanoscale superlattice of interacting quantum wires with a multicomponent electronic structure. We present first-principles quantum calculation of a universal superconducting dome where TcT_{c} amplification in multi-gap superconductivity is driven by the Fano-Feshbach resonance due to configuration interaction between open and closed pairing channels, i.e., between multiple gaps in the BCS regime, resonating with a single gap in the BCS-BEC crossover regime. In the proposed three dimensional (3D) phase diagram the critical temperature shows a superconducting dome where TcT_{c} is a function of two variables (i) the Lifshitz parameter (η\eta) measuring the separation of the chemical potential from the Lifshitz transition normalized by the inter-wires coupling and (ii) the effective electron phonon coupling (g) in the appearing new Fermi surface including phonon softening. The results will be of help for material design of room temperature superconductors at ambient pressure.

preprint: AIP/123-QED

I Introduction

I.1 Phenomenological overview

Pressurized sulfur hydride H3SH_{3}S, with TcT_{c}=203K203\ K at 162GPa162\ GPa [Drozdov et al., 2015], has reached in 2015 the record for the highest critical temperature, held before by cuprate perovskites since 1986 [Bednorz and Müller, 1988,Gao et al., 1994,Yamamoto et al., 2015]. This discovery has been followed by superconductivity in lanthanum hydrides with TcT_{c} above 260K260\ K [Somayazulu et al., 2019,Drozdov et al., 2019], in yttrium hydrides with Tc=243KT_{c}=243\ K [Troyan et al., 2021,Kong et al., 2021], and in a ternary carbonaceous sulfur hydride CSHxCSH_{x} [Snider et al., 2020] reaching room temperature. X-ray diffraction, using focused synchrotron radiation, has shown the crystalline Im3¯mIm\bar{3}m lattice symmetry of H3SH_{3}S above 103GPa103\ GPa [Einaga et al., 2016; Goncharov et al., 2017; Duan et al., 2017; Kruglov et al., 2017; Duan et al., 2017] and X-ray absorption spectroscopy has provided information on the local structure of yttrium hydrides [Purans et al., 2021].

Recent experimental results show an anomalous superconductivity phase [Troyan et al., 2021,Kong et al., 2021], while conventional superconductivity [Eliashberg, 1960,Dynes, 1972], considering only the pairing of the superconducting electrons via electron-phonon coupling (Cooper pairs) and a single-gap superconductivity paradigm, has been used to predict and to explain the high critical temperature in pressurized hydrides since the early days [Duan et al., 2014, 2015; Durajski and Szczęśniak, 2017; Gorkov and Kresin, 2018; Kostrzewa et al., 2020].

The old single-gap paradigm was found to be incompatible with band structure calculations of H3SH_{3}S in the pressure range where the critical temperature shows its maximum value, TcmaxT_{c\ max}. In fact band-structure calculations [Bianconi and Jarlborg, 2015a, b; Jarlborg and Bianconi, 2016], show that:

  1. (i)

    the applied pressure induces an increasing compressive lattice strain which pushes an incipient density of states (DOS) peak, due to a van Hove singularity (vHS), to higher energy until it crosses the Fermi level, [Bianconi and Jarlborg, 2015a] as confirmed by several authors [Quan and Pickett, 2016,Souza and Marsiglio, 2017];

  2. (ii)

    multiple Fermi surfaces coexist in different spots of the kk-space, [Bianconi and Jarlborg, 2015b];

  3. (iii)

    the Migdal approximation EFn{E_{Fn}} >> ω0{\hbar\omega_{0}} in the appearing nth Fermi-surface spot breaks down near the Lifshitz transition [Jarlborg and Bianconi, 2016];

  4. (iv)

    the anomalous pressure-dependent isotope coefficient [Jarlborg and Bianconi, 2016] strongly deviates from the single-band constant value predicted by the standard BCS theory.

The Fermi energy EFn{E_{Fn}} in the appearing new nth Fermi-surface spot at the Lifshitz transition and the energy width of the vHS singularity are of the order of the energy of the optical phonon ω0{\hbar\omega_{0}}=160meV160\ meV, observed by Capitani et al.[Capitani et al., 2017] in pressurized H3SH_{3}S infrared spectra. The latter show a Fano lineshape with the characteristic strong asymmetry indicating its interference with electronic degrees of freedom at the Fermi level [Fano, 1935,Fano, 1961].

The vHS at the Fermi energy, by using band-structure calculations, has been attributed to an electronic band of s orbitals originating from the network of hydrogen chains with short HHH-H hydrogen bonds [Jarlborg and Bianconi, 2016]. The lattice compressive strain, due to increasing pressure, induces the energy shift of the vHS. The latter crosses the chemical potential yielding a Lifshitz transition for the appearing of a new small Fermi-surface spot [Bianconi and Jarlborg, 2015a], while the other Fermi surfaces contribute to the featureless weak broad background of the density of states. The Lifshitz transition belongs to the class of electronic topological transitions [Lifshitz et al., 1960; Volovik, 2017; Volovik and Zhang, 2017; Volovik, 2018] for the appearing of a new Fermi surfaceof the 2.5 order for standard Fermi gases. These transitions become first order showing arrested phase separation for strongly interacting fermions [Kugel et al., 2008,Bianconi et al., 2015].

The critical temperature as a function of pressure in H3SH_{3}S and CSHxCSH_{x} is shown in Fig.1. The external pressure induces a compressive strain, shown in panel (a) of Fig.1. The strain is given by ϵ=100(aa0)/a0\epsilon=100(a-a_{0})/a_{0}, where a0a_{0} is the lattice constant at P=103GPaP=103\ GPa, where the Im3¯mIm\bar{3}m lattice symmetry appears because of a structural phase transition. The superconducting domedome over the strain range 0.5<ϵ<40.5<\epsilon<4 shows the maximum TcT_{c} at ϵ=2.5\epsilon=2.5. Panel (b) of Fig.1 shows the superconducting domedome observed in CSHxCSH_{x} [Snider et al., 2020]. Comparing Panels (a) and (b) we notice that the maximum TcT_{c} is higher in the superconducting dome of CSHxCSH_{x} while its width is smaller.

I.2 The Fano-Feshbach resonance in multi-gap superconductivity

The paradigm shift to multi-gap superconductivity including the key role of Majorana exchange interaction between different condensates [Bianconi, 2003,Vittorini-Orgeas and Bianconi, 2009,Palumbo, Marcelli, and Bianconi, 2016] has been proposed since 2015 [Bianconi and Jarlborg, 2015a]. The Bogoljubov formulation of superconductivity, beyond the attractive BCS force between two electrons via the exchange of a phonon, includes also the attractive Majorana or repulsive Heisenberg exchange interactions [Bogoljubov, Tolmachov, and Širkov, 1958] as in nuclear matter. In the latter the forces which are commonly assumed in the phenomenological proton-neutron Hamiltonian include

  1. (i)

    the Heisenberg exchange operator for particles which exhibit antisymmetric states, which interchanges both position and spin coordinates [Heisenberg, 1933];

  2. (ii)

    the Majorana exchange operator for particles which exhibit symmetric states, which interchanges the positions of the particles, leaving their spin directions unaffected [Majorana, 1933];

  3. (iii)

    the nuclear force, resulting from the exchange of mesons between neighboring nucleons (Yukawa type) [Yukawa, 1935].

The nuclear force has been called also the short range Wigner force, applied exclusively to non-exchange forces to account for the explanation of the large binding energy of He4 in comparison with the deuteron, as well as of the main features of neutron-proton scattering. The interplay of these forces in the many-body quantum physical description of nuclear matter has been the object of extended studies [Iachello, 2013; Feshbach and Iachello, 1974; Blatt and Weisskopf, 1979].

The proposed scenario of multi-gap superconductivity including exchange interactions near a Lifshitz transition in pressurized H3SH_{3}S [Bianconi and Jarlborg, 2015b] was previously proposed by Bianconi, Perali and Valletta (BPV) for other non BCS superconductors like hole-doped cuprates [Perali et al., 1996; Valletta et al., 1997; Bianconi et al., 1998; Bianconi, 2006; Perali et al., 2012], diborides [Bianconi et al., 2001a; Bianconi, 2003; Perali et al., 2004; Perali, Pieri, and Strinati, 2004; Bianconi, 2005], iron-doped superconductors [Innocenti et al., 2010a, b; Bianconi, 2012; Innocenti and Bianconi, 2013; Bianconi, 2013a], organics [Mazziotti et al., 2017; Pinto et al., 2020], metallic nanoscale multilayers with nodal lines where the spin-orbit interaction plays a key role in the TcT_{c} amplification [Mazziotti et al., 2021a], superconductivity [Bianconi et al., 2014] at the interface of oxide perovskites which can host also Majorana fermions [Mazziotti et al., 2018]. The BPV theory focuses on the quantum Fano-Feshbach resonance due to the configuration interaction between the open and the closed scattering channels [Bianconi, 2003,Vittorini-Orgeas and Bianconi, 2009,Palumbo, Marcelli, and Bianconi, 2016]. The Fano-Feshbach resonance was first proposed theoretically in atomic physics by Fano in 1935 [Fano, 1935,Fano, 1961] and extended by Feshbach in 1962 in the many-body physics of nuclear matter [Feshbach, 1962], where it is called shape resonance and it is described by the multi-channel optical model.

Refer to caption
Figure 1: Panel a: The superconducting domedome of H3SH_{3}S for P>120GPaP>120\ GPa with the critical temperature Tcmax=203KT_{c\ max}=203K at its top at Popt=162GPaP_{opt}=162\ GPa with half width of about 43GPa43\ GPa. Panel b: The superconducting domedome of CSHxCSH_{x} for P>220GPaP>220\ GPa with Tcmax=287KT_{c\ max}=287K at Popt=265GPaP_{opt}=265\ GPa

In the quantum theory of the many-body systems, made of different electronic components, the Fano-Feshbach resonance appears when the Fermi wavelength of one of the components is of the order of the system size as in nuclear matter [Feshbach, 1983,Feshbach, 2014] and in condensed matter at the nanoscale [Miroshnichenko, Flach, and Kivshar, 2010]. Shape resonances have been found in the final states of X-ray absorption near edge structure where the photoelectron wavelength is of the order of interatomic distance and the electronic multiple scattering resonance is degenerate with the continuum [Bianconi et al., 1978; Bianconi, 1980; Bianconi et al., 1982].

The exchange interaction between condensates was included in the theories for overlapping bands by Suhl, Matthias, and Walker (SMW) [Suhl, Matthias, and Walker, 1959], Moskalenko [Moskalenko, 1959] and Kondo [Kondo, 1963], even though they assumed, in a first approximation, that all intraband pairing channels in each of the nn bands were in the BCS regime with (EFnω0E_{Fn}\gg\omega_{0}). Furthermore the exchange term for interband pair transfer was assumed to be a constant parameter with no energy or momentum dependence. Therefore the above theories of overlapping bands could not include Fano-Feshbach resonances. Indeed, the Fano-Feshbach resonance in the Bogoliubov superconductivity theory of multi-gap superconductors is due to the configuration interaction between different pairing channels in different Fermi surfaces [Mazziotti et al., 2021b] with exchange of pairs between the first condensate in the BCS regime and second condensate in the BCS-BEC crossover regime. In the BCS-BEC regime the momentum and energy dependence of the exchange interaction between different coexisting gaps plays a key role, while it is neglected in the anisotropic Eliashberg theory of multi-gap superconductors. On the contrary, in the BPV theory [Perali et al., 1996], the Fano-Feshbach resonance between a firstfirst pairing channel (called closed channel) in the BCS-BEC crossover regime, and the open pairing channels (called open channels) in other large Fermi surfaces in the BCS regime has been calculated from the overlap of the wave-functions of the electron pairs in different bands. The latter are determined by the subtle overlap of the wave-functions of pairs in superlattices of interacting 1D or 2D units. In ultracold fermion gases the Fano-Feshbach resonance has been applied in 2004 to generate unconventional fermion superfluids with a very large ratio of Tc/TFT_{c}/T_{F} [Zwierlein et al., 2004,Zhang et al., 2004]. The quantum amplification mechanism at Fano-Feshbach resonance near a Lifshitz transition is generated by the quantum interference of pairing between:

  1. (i)

    electrons in the new appearing small Fermi surface with low Fermi energy and Fermi wavelength λF\lambda_{F} of the order of the system size;

  2. (ii)

    electrons in other Fermi surfaces with very high Fermi energies and very short Fermi wavelength λF\lambda_{F}.

At optimum TcT_{c} in the closed pairing channel, λF\lambda_{F} is larger but close to the superconducting coherence length kFξ010k_{F}\xi_{0}\sim 10 [Uemura et al., 1989,Pistolesi and Strinati Calvanese, 1994].

I.3 Choice of the model

Here we propose that the experimental superconducting domedome, given by the curves of the critical temperature TcT_{c} as a function of pressure P in Fig.1 in sulfur hydrides, is the smoking gun of the Fano-Feshbach resonance between pairing channels driven by the variable lattice strain [Perali et al., 1996,Mazziotti et al., 2017,Agrestini et al., 2003] which tunes the chemical potential at a topological Lifshitz transition.

The Fano-Feshbach resonance, in multi-gap superconductivity in H3SH_{3}S, is supported by the unusual behavior of the isotope coefficient. Indeed, the isotope coefficient decreases from 2.372.37 to 0.310.31 in the range going from the threshold to the top of the superconducting domedome [Jarlborg and Bianconi, 2016; Szczesniak and Durajski, 2017, 2018; Drozdov et al., 2019], deviating markedly from the value 0.50.5, predicted by the single-band BCS theory. A similar anomalous behavior of the isotope coefficient has been found in the superconducting domedome of cuprate perovskites [Perali et al., 1997],[Bianconi et al., 1998],[Perali et al., 2012,Innocenti and Bianconi, 2013].

A nanoscale heterostructure is expected to appear in compounds made by combining several chemical elements which leads to competing orders of electronic degrees of freedom. In pressurized hydrides the nanoscale heterogeneity is determined by the local lattice structure controlled by the effects of the lattice strain resulting from the interplay between lattice misfit-strain (or chemical pre-compression) and the external pressure. These heterostructures are a particular case of a supersolid stripes crystal [Bianconi, Innocenti, and Campi, 2013] called superstripes [Bianconi, 2013a, b], which can be realized with optical lattices in ultracold gases [Masella et al., 2019].

Refer to caption
Figure 2: The upper panel shows the unit cell of pressurized H3SH_{3}S crystal structure at 150GPa150\ GPa with cubic Im3¯mIm\bar{3}m lattice. The central and lower panels show that at mesoscale H3SH_{3}S appears to be made of stacks of H2H_{2} layers (small green dots) in the [101] plane intercalated by HSHS layers where SS is indicated by large yellow dots. The central projection of the H3SH_{3}S crystal shows that the H2H_{2} layers are made of atomic hydrogen chains with the shortest H-H metallic bonds in the [100][100] direction.

The superconductivity in the superstripes phase with coexisting different localised and delocalised electronic components moving in complex nanoscale heterostructures of low dimensional (quasi one-dimensional 1D) structural units (called chains or stripes or ladders) has been found in:

  1. (i)

    A15 intermetallics which have held the record for the highest superconducting Tc=23.2KT_{c}=23.2\ K from 1973 to 1986. A15 intermetallics like Nb3GeNb_{3}Ge and Nb3SnNb_{3}Sn have the same average crystal symmetry Im3¯mIm\bar{3}m as H3SH_{3}S [Mazziotti et al., 2021b] and show complex textures made of a metallic 3D network of interacting 1D metallic Nb chains [Testardi, 1975].

  2. (ii)

    hole doped cuprate perovskites where 2D networks of extrinsic stripes with different local lattice distortions [Bianconi et al., 1991, 1996a, 1996b] appear at nanoscale in the CuO2CuO_{2} atomic layers facilitated by the polymorphism of perovskite structures. These form metamorphic lattice stripes in mismatched material systems [Gavrichkov et al., 2019] whose mismatch is tuned by the lattice misfit strain [Agrestini et al., 2003,Bianconi et al., 2000,Di Castro et al., 2000].

  3. (iii)

    superconducting organics like doped p-terphenyl [Pinto et al., 2020], where 1D-wires of short hydrogen bonds have been observed by X-ray diffraction [Barba et al., 2018] and it was proposed that the high-TcT_{c} is driven by the Fano-Feshbach resonance in the nanoscale superlattice of quantum wires [Mazziotti et al., 2017].

Superlattices of two-dimensional (2D) metallic quantum wells at nanoscale have been found in:

  1. (i)

    diborides made of stacks of boron layers intercalated by magnesium [Bauer et al., 2001; Agrestini et al., 2001; Bianconi et al., 2001b; Agrestini et al., 2004; Di Castro et al., 2002];

  2. (ii)

    iron-based perovskite superconductors, iso-structural with electron doped cuprates, [Ricci et al., 2009,Ricci et al., 2010], which are made of stacks of iron atomic layers and the tuning of the chemical potential near the Lifshitz transition has been clearly seen in ARPES experiments [Bianconi, 2013a,Kordyuk, 2018,Pustovit and Kordyuk, 2016].

In this work we present the theoretical prediction of the superconducting dome for room-temperature superconductivity in pressurized hydrides due to a Fano-Feshbach resonance near a Lifshitz transition in the frame of the multi-gap superconductivity scenario discussed recently by several authors [Guidini et al., 2016; Cariglia et al., 2016; Doria, Cariglia, and Perali, 2016; Bussmann-Holder et al., 2016; Valentinis, Van Der Marel, and Berthod, 2016; Bussmann-Holder et al., 2017, 2019; Chubukov and Mozyrsky, 2018; Salasnich et al., 2019; Kagan and Bianconi, 2019; Tajima, Perali, and Pieri, 2020; Vargas-Paredes et al., 2020].

A key feature of our approach is the inclusion of the exchange integrals, obtained by the overlap of the wave-functions of electrons in different Fermi surfaces. We evaluate them by solving the Schrödinger equation for a lattice heterostructure including the renormalisation of the chemical potential at the Lifshitz transition with the opening of a new superconducting gap, controlled by the constraint of the number density equation. The results provide a significant step in understanding room-temperature superconductivity and the physical origin of the superconducting dome. Moreover, the results indicate a road map for the material design of artificial mesoscopic heterostructures made of nanoscale quantum wires which can be used by material scientists to synthesize new room-temperature superconductors at ambient pressure.

II The van Hove Singularity in the superstripes phase

Refer to caption
Figure 3: Bottom panels: total Density of States (DOS) for the case A, panel a (for the case B, panel b) shown by the thick solid blue (black) line as a function of the Lifshitz parameter η\eta of the superlattice of quantum wires with weak inter-wire interaction giving the transversal dispersion ΔE=274meV\Delta E=274\ meV (145meV145\ meV). The figure shows also the high partial DOS curves at the van Hove singularity due to the upper third subband (red line) with small Fermi energy and the low partial DOS curves due to first and second subbands (blue dashed lines) with high Fermi energies. The top panel c shows the Lifshitz topological transition in the Fermi surfaces due to the third subband, shown with a solid red line, where the Fermi surfaces for the first and second subbands are indicated by blue and orange lines. The appearing Fermi surface due to the third subband changes at three different values of η\eta indicated in panel a: η1\eta_{1} corresponds to the appearing of a small tubular Fermi surface in the kx,kzk_{x},k_{z} plane; η2\eta_{2} is the energy where the size of the tubular Fermi surface becomes large and it is close to the Lifshitz transition for neck disrupting (or opening a neck) where its topology changes from 2D for η2\eta_{2} to 1D for η3\eta_{3} [Bianconi and Jarlborg, 2015a].
Refer to caption
Figure 4: The softening of the phonon energy ω~0\tilde{\omega}_{0} according to the Migdal theory, as a function of the intraband electron-phonon coupling gg in the small Fermi surface spot appearing at the Lifshitz transition for the case A (blue line) and for the case B (red line)

We propose that in agreement with [Bianconi, 1996,Bianconi, 2001,Bianconi, 1994] a heterostructure at atomic limit made of superconducting quantum wires running in the xx-direction intercalated by spacers of thickness WW with periodicity d=W+Ld=W+L (in the zz-direction). The metallic wires are separated by spacers which generate a potential barrier of amplitude V [Bianconi, 1996, 2001, 1994], where the parameters values are chosen in order to capture the main features of the DOS at the van Hove singularity near the Fermi level of H3SH_{3}S. The lineshape of the DOS peak shows the features of a high-order anisotropic van Hove singularity typical of a supermetal [Isobe and Fu, 2019], In this superstripes phase the superconductivity is calculated using the BPV approach first proposed for striped cuprate perovskites [Perali et al., 1996, 1997; Bianconi et al., 1997; Valletta et al., 1997; Bianconi et al., 1998; Bianconi, 2006; Perali et al., 2012].

We present numerical calculations of the multi-gap superconductivity domedome of the critical temperature as a function of pressure where we change both i)i) the proximity of the chemical potential to the Lifshitz transition, ii)ii) the electron-phonon coupling for electrons in the upper subband and the iii)iii) the phonon softening for increasing electron-phonon coupling g in the disappearing Fermi surface.

The proximity to the Lifshitz transition is measured by the Lifshitz parameter η\eta given by the energy difference between the chemical potential and the energy ELE_{L} of the topological Lifshitz transition, which is the band-edge energy of the highest energy subband, normalized to the transversal energy dispersion, ΔE\Delta E, between the 1D metallic chains

η=μELΔE.\eta=\frac{\mu-E_{L}}{\Delta E}.

The applied external pressure induces the variation of either i)i) of the Lifshitz parameter η\eta and ii)ii) of the electron-phonon coupling joint with phonon energy softening of the particular phonon mode coupled with the electrons in the small Fermi surface spot in the new appearing subband. In the heterostructure of quantum wires the electrons along the xx-direction are free, while along the zz-direction they are subjected to a periodic potential. Hence, the eigenfunctions ψnkz(z)\psi_{nk_{z}}(z) and the eigenvalues En(kz)E_{n}(k_{z}), along the confinement direction, can be computed only numerically by solving a corresponding Kronig-Penney model. The solution of the eigenvalues equation gives the electronic dispersion for the nn subbands. Indeed, in the heterostructure of quantum wires, the quantum-size effects give a multiband electronic structure where the subband with higher energy shows a two-dimensional behavior due to hopping between 1D-chains. For the numerical calculation of the superconducting dome of H3SH_{3}S we start with the evaluation of the DOS peak in [Jarlborg and Bianconi, 2016] at EFE_{F} by using a model of 1D-chains corresponding to the chains with short H-H bonds, as shown in Fig.2. We have designed an artificial nanoscale heterostructure at atomic limit made of quantum wires of width L=0.85L=0.85 nm, spacers of width W=0.55W=0.55 nm separated by a potential barrier V=4.16 eV which reproduces the van Hove singularity in a range of 500 meV around the Fermi energy H3SH_{3}S i.e., within the energy cut-off of the pairing interaction relevant for the emergence of superconductivity. We consider the models A and B for the heterostructure characterized by different coupling between the quantum wires i.e., with different transversal dispersion ΔE\Delta E obtained by changing the effective mass in the spacers and in the wires. The model A of the superlattice of wires is characterized by ΔE\Delta E= 274meV274\ meV transversal dispersion as it was found in H3SH_{3}S. The model B is characterized by a smaller transversal dispersion ΔE=145meV\Delta E=145\ meV, obtained by increasing the effective mass in the barrier. A smaller dispersion is expected to give a sharper superconducting domedome and a higher maximum critical temperature close to room temperature. The DOS peak and the partial DOS for the model A and model B as a function of the Lifshitz parameter are plotted in panel (a) and panel (b) of Fig.3 respectively. Panel (c) of Fig.3 shows the Lifshitz transition for the appearing of a new Fermi surface, called of type (I), tuning the chemical potential near the band-edge (η1\eta_{1}) of the subband with a critical point where a new 2D Fermi surface spot appears. The second type of Lifshitz transition (type II) occurs at the opening of a neck in the small Fermi surface with the appearing of a singular nodal point which gives the sharp DOS maximum at η2\eta_{2} (Fig.3 panel (c)) at the crossover between the 2D and 1D topology. The nearly flat portion of the DOS between (type I) and (type II) Lifshitz transitions in Fig.3 correspond with the regime where a small 2D Fermi surface with a low Fermi energy appears. While in previous theoretical descriptions of the Fano-Feshbach resonance near the Lifshitz transition the electron-phonon coupling gnng_{nn} was assumed to be constant, in this work we take into account that the external pressure changes either the Lifshitz parameter (η\eta) and the electron-phonon coupling gnng_{nn} in the appearing Fermi surface in the upper subband and the renormalized phonon energy ω~0\tilde{\omega}_{0} shows the softening according to the Migdal relation:

ω~0=ω012×maxngnn.\tilde{\omega}_{0}=\omega_{0}\sqrt{1-{\color[rgb]{0,0,0}2\times\max_{n}g_{nn}}}. (1)

The relation contains the coupling constant for the metal forming the superconducting layers for g<0.5g<0.5, and it is used here to qualitatively estimate the effect of the coupling constant on the phonon frequency in the appearing Fermi surface as it is shown in Fig.4. In our theory, the variable ω~0\tilde{\omega}_{0} is also the cut-off energy of the pairing interaction in the Bogoliubov gap equation which changes with η\eta. We have fixed, for the case A, ω0=330meV\omega_{0}=330\ meV in order to reproduce with moderate intraband electron phonon coupling, 0.3<g<0.330.3<g<0.33, the experimental phonon frequency, ω~0=160meV\tilde{\omega}_{0}=160\ meV [Capitani et al., 2017], measured in pressurized H3SH_{3}S at 150GPa150\ GPa For the case B we have fixed ω0=225meV\omega_{0}=225\ meV, in order to get ω~0=ΔE\tilde{\omega}_{0}=\Delta E = 145meV145\ meV with moderate intraband electron phonon coupling g=0.25g=0.25.

In the case of organic superconductors [Mazziotti et al., 2017] it has been shown that the amplification of the critical temperature in heterostructures of quantum wires and a narrow superconducting dome occurs where the coupling in the appearing new nth Fermi surface is larger than the in other Fermi surfaces and the interband coupling is small. In our model gnng_{nn^{\prime}} is the superconducting dimensionless coupling constant for the three-band system which has a matrix structure that depends on the band indices nn and nn^{\prime}.

In this work we confirm previous results [Mazziotti et al., 2017]: in fact the superconducting dome with a sharp drop of TcT_{c} at both sides of the maximum with a stronger Fano-Feshbach anti-resonance is generated by a weak intra-band coupling for the Cooper pairing channel gnng_{nn} and weak inter-band exchange channels gnng_{nn^{\prime}}. The Fano-Feshbach resonance increases the maximum value of TcT_{c} at the top of the domedome increasing the intra-band coupling for the Cooper pairing channel gnng_{nn} in the new appearing or disappearing Fermi surface. Moreover the maximum of the critical temperature is expected to increase where the phonon energy giving the energy cut off for the pairing processes is of the same order as the hopping energy between the wires ΔE=ω~\Delta E=\tilde{\omega}.

When the Lifshitz parameter is tuned between the band-edge and the van Hove singularity, a new Fermi surface appears with a very small number density of electrons in the strong coupling limit. The condensates in the other Fermi surfaces (first and second subband) have a very high Fermi energy and therefore are in the adiabatic regime and coexist with a third condensate in the small Fermi surface where the classical BCS approximations are violated. In the models (A) and (B) for 0<η\eta<1 a new closed 2D Fermi surface appears as shown in band structure calculations [Bianconi and Jarlborg, 2015a] for H3SH_{3}S around 160 GPa where the maximum critical temperature is observed at a top of a domedome. For this heterostructure we assume that quantum size effects are not negligible and the electron hopping in the transverse direction is finite so that the quantum wires can be considered to be interacting wires in the metallic phase while very weakly interacting wires are in the localization limit. This is reflected in the spectrum that appears to split into nn subbands characterized by quantized values of the transverse moment that depends on the band index and the dimension of the wires.

In the heterostructure of quantum wires the electrons along the xx-direction are free, while along the zz direction they are subjected to a periodic potential V(z)V(z):

V(z)=Vθ(W/2|mλpz|).V(z)=V\sum^{\infty}_{-\infty}\theta(W/2-|m\lambda_{p}-z|). (2)

In the periodic potential we assume that the full single-particle wave-function can be written as

ψn,𝐤,α(𝐫)=1LxLzeikxxψnkz(z)𝝌α,\psi_{n,\mathbf{k},\alpha}(\mathbf{r})=\frac{1}{\sqrt{L_{x}L_{z}}}e^{ik_{x}x}\psi_{nk_{z}}(z)\bm{\chi}_{\alpha}, (3)

where LxL_{x} and LzL_{z} are the spatial dimensions of the system, nn is the band index, 𝐤=(kx,kz)\mathbf{k}=(k_{x},k_{z}) is the wavevector, and 𝝌α\bm{\chi}_{\alpha} is the spinor part with spin α=\alpha=\uparrow or \downarrow. The corresponding energy eigenvalues, independents from the spin, are given by

εn(𝐤)=22mxkx2+En(kz).\varepsilon_{n}(\mathbf{k})=\frac{\hbar^{2}}{2m_{x}}k^{2}_{x}+E_{n}(k_{z}). (4)

The eigenfunctions ψnkz(z)\psi_{nk_{z}}(z) and the eigenvalues En(kz)E_{n}(k_{z}) are computed numerically by solving a corresponding Kronig-Penney model. The solution of the eigenvalues equation gives the electronic dispersion for the nn subbands.

III multi-gap superconductivity beyond BCS

In the multi-gap superconducting scenario the exchange integral for pairs of electrons in different bands plays a key role for the TcT_{c} amplification while it is neglected in the single-band BCS theory.

The pairing interaction is assumed to be originated from an electron-electron contact interaction with a cut-off equal to the renormalized phonon energy ω~0\tilde{\omega}_{0} . The pairing interaction takes then a generalized BCS form

U𝐤𝐤nn=U~kzkznnθ(ω~0|ξn,𝐤|)θ(ω~0|ξn,𝐤|)U_{{\bf k}{\bf k^{\prime}}}^{nn^{\prime}}=\tilde{U}_{k_{z}k^{\prime}_{z}}^{nn^{\prime}}\theta(\tilde{\omega}_{0}-|\xi_{n,{\bf k}}|)\theta(\tilde{\omega}_{0}-|\xi_{n^{\prime},{\bf k^{\prime}}}|) (5)

where ξn,𝐤=εn(𝐤)μ\xi_{n,{\bf k}}=\varepsilon_{n}(\mathbf{k})-\mu. The U~kzkznn\tilde{U}_{k_{z}k^{\prime}_{z}}^{nn^{\prime}} coupling constants, which in the original BCS model are structureless, originate from the matrix elements between exact eigenstates of the superlattice, and depend on the wave vectors kzk_{z} and kzk_{z}^{\prime} in the superlattice direction as well as on the band indices nn and nn^{\prime}. This induces a structure in the k-dependent interband coupling interaction for the electrons that determines the quantum interference between electron pairs wave functions in different subbands or minibands of the superlattice. The generalized couplings can be expressed as

U~kzkznn=U0nnIkzkznn,\tilde{U}_{k_{z}k^{\prime}_{z}}^{nn^{\prime}}=-{\color[rgb]{0,0,0}U_{0}^{nn^{\prime}}}\ I_{k_{z}k^{\prime}_{z}}^{nn^{\prime}},

where U0nn-{\color[rgb]{0,0,0}U_{0}^{nn^{\prime}}} is the original attractive contact interaction, that we allow to have a dependence from the band index of the electron pairs, and IkzkznnI_{k_{z}k^{\prime}_{z}}^{nn^{\prime}} is the pair superposition integral, calculated considering the interference between electronic wave functions in different subbands [Innocenti et al., 2010a].

Ikzkznn=1LxLz2ψnkz(z)ψnkz(z)ψnkz(z)ψnkz(z)𝑑z.I_{k_{z}k^{\prime}_{z}}^{nn^{\prime}}=\frac{1}{L_{x}L^{2}_{z}}\int\psi_{nk_{z}}^{*}(z)\psi_{n-k_{z}}^{*}(z)\psi_{n^{\prime}k^{\prime}_{z}}(z)\psi_{n^{\prime}-k^{\prime}_{z}}(z)dz.

Notice that for vanishing VV, the amplitude of the periodic potential, the overlap integrals would reduce to the standard BCS form Ikzkznn=(LxLz)1I_{k_{z}k^{\prime}_{z}}^{nn^{\prime}}=(L_{x}L_{z})^{-1} independent of the kzk_{z} and kzk_{z}^{\prime} wave vectors as well. We emphasize that the exchange interaction is not constant but depends not only on the wave vector along z but also on the band index, therefore it has a matrix structure. For later reference and to compare with the homogeneous case, it is useful to introduce the standard dimensionless coupling constant gnn=U0nnN0g_{nn^{\prime}}=U_{0}^{nn^{\prime}}N_{0}, where N0N_{0} is the two-dimensional density of states. The non diagonal terms (nnnn^{\prime}) of the superposition integral are calculated for model A and are shown in Fig.5.

Refer to caption
Figure 5: Panel A: histogram of the superposition integral of equation (5). The numbers indicate the different elements of the matrix of the intraband pairings Ikzkzn,nI_{{k}_{z}{k}^{\prime}_{z}}^{n,n^{\prime}} and the interband couplings Ikzkzn,nI_{{k}_{z}{k}^{\prime}_{z}}^{n,n^{\prime}}. Panel B: the matrix elements of the exchange integral as a function of the wavevectors in the direction of the confinement potential. The colors correspond to those of the histogram.

The self-consistent equation for the superconducting gap at zero temperature can then be written as

Δnkz=12n,𝐤U𝐤𝐤nnΔnkz(εn(𝐤)μ)2+|Δnkz|2,\Delta_{nk_{z}}=-\frac{1}{2}\sum_{n^{\prime},\mathbf{k}^{\prime}}\frac{U^{nn^{\prime}}_{{\bf k}{\bf k^{\prime}}}\Delta_{n^{\prime}k^{\prime}_{z}}}{\sqrt{(\varepsilon_{n^{\prime}}(\mathbf{k^{\prime}})-\mu)^{2}+|\Delta_{n^{\prime}k^{\prime}_{z}}|^{2}}}, (6)

In order to take into account the renormalisation of the chemical potential and charge densities in each subband when a new superconducting gap appears in a single subband, the joint Bogoliubov gap equation and the charge density equation have been solved where the charge density ρ\rho and the chemical potential in the superconducting phase are related by

ρ=1LxLzn𝐤(1εn(𝐤)μ(εn(𝐤)μ)2+Δn,kz2)θ(μεn(𝐤)),\rho=\frac{1}{L_{x}L_{z}}\sum_{n\mathbf{k}}\bigg{(}1-\frac{\varepsilon_{n}(\mathbf{k})-\mu}{\sqrt{(\varepsilon_{n}(\mathbf{k})-\mu)^{2}+\Delta^{2}_{n,k_{z}}}}\bigg{)}\theta(\mu-\varepsilon_{n}(\mathbf{k})), (7)

The joint solution of the gap equation (3) and the density equation (4) is essential in order to correctly describe the multi-gap superconductivity near the Lifshitz transition where the gap in the upper subband approaches to the Bose-Einstein condensation.

The superconducting critical temperature is calculated by iteratively solving the linearized equation

Δn𝐤=12n𝐤UkzkznnΔnkztanh((εn(𝐤)μ)2Tc)(εn(𝐤)μ)\Delta_{n\mathbf{k}}=-\frac{1}{2}\sum_{n^{\prime}\mathbf{k}^{\prime}}U_{{k}_{z}{k^{\prime}}_{z}}^{nn^{\prime}}\Delta_{n^{\prime}k^{\prime}_{z}}\frac{\tanh\left(\frac{(\varepsilon_{n^{\prime}}(\mathbf{k}^{\prime})-\mu)}{2T_{c}}\right)}{(\varepsilon_{n^{\prime}}(\mathbf{k}^{\prime})-\mu)} (8)

until the vanishing solution is reached with increasing temperature.

Here we present a case of Fano-Feshbach resonant superconductivity giving a superconducting domedome where the top of the domedome reaches the high temperature range 200<Tcmax<300K200<T_{c}\ max<300K of pressurized hydrides much larger than Tcmax=123KT_{c}\ max=123K calculated in a previous work for cuprates and organics [Mazziotti et al., 2017]. This result is obtained by the resonance regime by increasing gaps anisotropy where the two gaps differ by a sizable factor in the range 2.9-3.9, at the top of the dome where the coupling strength in a small Fermi surface spot is in the range 0.3<g<0.42 and the phonon energy scale determines not only a large prefactor for the critical temperature, but it also induces a large width of the resonance.

Here, following Ref.[Mazziotti et al., 2017], the superconducting dome is generated by considering the case where the first and the second subband are in a weak coupling regime because the Fermi level is very far from the band edge. Therefore in our model we fixed the values of U0nnU_{0}^{nn^{\prime}} in order to have the following values for the dimensionless coupling constants: g11=g22=0.10g_{11}=g_{22}=0.10. On the contrary the coupling in the third subband g33=gg_{33}=g is considered to be variable because the Fermi level is tuned around the band edge. In fact we expect that the electron-phonon coupling for the upper subband should be enhanced because of a Kohn anomaly or because the interplay with the formation of a charge density wave (CDW) in a narrow momentum region around the CDW wave vector. In parallel, we vary the cut-off energy ω~0\tilde{\omega}_{0} according to the Migdal relation [Migdal, 1958].

Refer to caption
Figure 6: Panel 𝐀𝟏\mathbf{A_{1}}: gap ratio as a function of the Lifshitz parameter for g=1/4g=1/4 (case (A)) for the second and third subband. Panel 𝐁𝟏\mathbf{B_{1}}: isotope coefficient as a function of the Lifshitz parameter for g=1/4g=1/4 (case (A)). Panel 𝐀𝟐\mathbf{A_{2}}: gap ratio as a function of the Lifshitz parameter for g=1/3g=1/3 (case (B)) for the second and third subband. Panel 𝐁𝟐\mathbf{B_{2}}: isotope coefficient as a function of the Lifshitz parameter for g=1/3g=1/3 (case (B)).
Refer to caption
Figure 7: The critical temperature as a function of the ratio between the gap in the third subband and the gap in the second subband. Panel A represents the trend for the case (A), while panel B represents the case (B). It can be noted that in the range 2.6<Δ3/Δ2<2.92.6<\Delta_{3}/\Delta_{2}<2.9 (for the case (A)) or 3.5<Δ3/Δ2<3.93.5<\Delta_{3}/\Delta_{2}<3.9 (for the case (B)) the critical temperature increases as the anisotropy between the gaps increases (blue arrows) until it reaches a maximum value when Δ3/Δ2\Delta_{3}/\Delta_{2} is maximum, from this point on then the TcT_{c} decreases almost exponentially as the ratio between the gaps decreases (red arrow). The blue circle represents the point where TcT_{c} is maximum, the red circle the point of intersection of the two opposite trends.
Refer to caption
Figure 8: Critical temperature as a function of the Lifshitz parameter (horizontal axis at the bottom) and as a function of the energy with respect to the band-edge (horizontal axis at the top) as the electron-phonon coupling varies, for two different models: case A (left panels) and case B (right panels). The curves in the bottom panels are plotted in a linear scale to show the evolution of the superconducting dome and in a semi-log scale in top panels to show the suppression of the critical temperature due to the Fano antiresonance of the left side of the dome. This plot shows that the critical temperature TcT_{c} for case A reaches a maximum value of 250K250\ K, while for case B it reaches a maximum value of 330K330\ K.

In Fig.6 the panel A1 (B1) shows the values of the gap ratio, 2Δ/Tc2\Delta/T_{c}, for the second and third subband for the A (B) case as a function of the Lifshitz parameter η\eta. The panel A2 (B2) in Fig.6 shows the trend of the isotope coefficient for the A (B) case as a function of the Lifshitz parameter η\eta. All these graphs were obtained at a fixed coupling value equal to g=1/4g=1/4 for the case (A) and at g=1/3g=1/3 for the case (B). At the Lifshitz transition for the appearing of a new Fermi surface, η=0\eta=0, the value of the gap ratio 2Δ/Tc2\Delta/T_{c} is close to the predicted value the BCS theory (2Δ/Tc=3.52\Delta/T_{c}=3.5) but for 0.5<η<1=00.5<\eta<1=0 we see strong deviations from the BCS single gap prediction. In fact, 2Δ2/Tc2\Delta_{2}/T_{c} reaches a very small value, between 0 and 1, while 2Δ3/Tc2\Delta_{3}/T_{c} remains approximately constant at the BCS value. A similar scenario was observed in magnesium diboride [Innocenti et al., 2010a] due to the exchange integral for pairs transfer between the second and third subbands.
While the BCS theory predicts that the isotope coefficient should be constant close to the value of 0.5 we see that the isotope coefficient shows a strong maximum of the Lifshitz transition η=0\eta=0, and a minimum at 0.5<η<1=00.5<\eta<1=0 near the topological Lifshitz transition for opening a neck in the Fermi surface of the third subband, These theoretical predictions are in agreement with the experiments showing that the isotope coefficient shows an anomalous trend as a function of pressure in pressurized sulfur hydrides [Jarlborg and Bianconi, 2016]. I

In Fig.7 we show the trend of the critical temperature as a function of the ratio between the gap in the third subband and the gap in the second subband for the case A in panel A and for the case B in panel B. The results clearly show that the maximum critical temperature is reached with the highest anisotropy between the gaps. In fact the graph shows that the maximum of TcT_{c} is reached when the ratio Δ3/Δ2\Delta_{3}/\Delta_{2} is maximum. This figure shows clearly that room-temperature superconductivity is reached by increasing electron-phonon coupling in the a small Fermi surface spot pushing up the gap in the appearing Fermi surface due to the third subband Δ3\Delta_{3} while the gap Δ2\Delta_{2} in the second Fermi surface with large Fermi energy remains small because the electron-phonon coupling remains small. These results show the predicted effect of Fano-Feshbach resonance driven by the exchange interaction between closed (strong) pairing channels in the third subband and open (weak) pairing channels in the second subband

IV Superconducting dome

In Fig.8 we plot the critical temperature as a function of the Lifshitz parameter in both semi-logarithmic and linear scales for variable values of the electron-phonon (e-ph) coupling in the upper subband. In the linear scale we see a variable superconducting dome where TcmaxT_{c\ max} increases with gg increasing up to g=0.4g=0.4 and it decreases in the range 0.4<g<0.50.4<g<0.5 because the phonon softening goes to zero at g=0.5g=0.5. In the case (A) the maximum value of the critical temperature is 250K250\ K, therefore it explain the superconductivity in H3SH_{3}S. While the maximum TcT_{c} in case (B) reaches 330K330\ K showing the possibility of room-temperature superconductivity. The plots in semi-logarithmic scale show the typical form of the Fano-Feshbach anti-resonance which becomes more relevant as gg increases. Fig.9 shows the variation of the critical temperature TcT_{c} at constant η\eta and the variable electron-phonon coupling gg for the A case. In the anti-resonant regime 1<η<0-1<\eta<0 we observe a clear feature of the Fano-Feshbach resonance. In fact, at the low energy side of the Fano-Feshbach resonance between closed and open channels, the negative interference gives the observed TcT_{c} minimum appearing at η=0.34\eta=-0.34 where the critical temperature decreases with increasing e-ph coupling, on the contrary for η>0\eta>0 TcT_{c} increases with increasing e-ph coupling up to g=0.4g=0.4.

Refer to caption
Figure 9: The critical temperature TcT_{c} as a function of the electron-phonon coupling gg at fixed different Lifshitz parameters η\eta for the case A. The curves Tc(g)T_{c}(g) in the lower part of the figure show the case with fixed η\eta in the anti-resonant regime 1<η<0-1<\eta<0 where the critical temperature decreasesdecreases by increasing the electron phonon coupling g reaching a minimum at η=0.34\eta=-0.34. The upper part of the figure shows the cases for η>0\eta>0 in the resonant regime, where the critical temperature increasesincreases by increasing g in the range 0.1<g<0.40.1<g<0.4 with the temperature scale in the right side.
Refer to caption
Figure 10: Calculated superconducting dome for H3SH_{3}S simulated using the proposed A model. The critical temperature is plotted in a color plot from blue (Tc=0KT_{c}=0K) to red (Tc=250KT_{c}=250K) as a function of two variables controlling the pairing strength in the new appearing small Fermi surface above the band edge of the third upper subband: (i) the Lifshitz parameter η\eta measuring the normalized Fermi energy EF3E_{F3} in the range 0<η<20<\eta<2 and (ii) the reduced Dynes [Dynes, 1972] electron-phonon coupling gg=(λ\lambda/(1+λ))\lambda)) with lambda equal to the bare pair coupling.

Fig.10 shows the critical temperature TcT_{c}(g,η\eta) as a function of two variables: (i)(i) the e-ph coupling in the third subband (g), where gg is the reduced Allen-Dynes electron-phonon coupling (λ\lambda/(1+λ\lambda)) [Dynes, 1972] and (ii)(ii) the Lifshitz energy parameter (η\eta). The critical temperature TcT_{c} is calculated by the BPV approach including the superconducting shape resonance between multiple gaps. The maximum TcT_{c} of the dome occurs in the (η,g)(\eta,g) plane at the point (1,0.4)(1,0.4) i.e.,i.e., at the Lifshitz transition for neck disrupting, at η\eta=1, which is associated with a transition of the topology of the small appearing Fermi surface from 1D at higher energy to 2D topology at lower energy. The universal superconductive dome obtained in this figure is needed to understand the experimental dome observed in the experimental curves of the critical pressure versus pressure TcT_{c}(P) of sulfur hydrides. In fact the external pressure induces a joint variation of both the energy position of the chemical potential with respect to the band-edge (the Lifshitz parameter η\eta ) as well as the electron-phonon coupling gg in the upper subband along a line in the (η\eta,gg) plane. The variable electron-phonon coupling is associated with the softening of the phonon mode energy coupled with electrons in the upper subband according to the Migdal relation. Therefore the experimental curve of TcT_{c} vs pressure shown in Fig.1 for a particular pressurized hydride is determined by different cuts of the universal superconductive dome determined by the particular pathway in the (η\eta,gg) plane driven by variable pressure.

Refer to caption
Figure 11: 3D color plot of TcT_{c} as a function of two variables (η,g\eta,g) calculated using the BPV approach for the heterostructure B, proposed here as the B model for pressurized CSHxCSH_{x}. The critical temperature increases from blue (Tc=0KT_{c}=0\ K) to red (Tc=300KT_{c}=300\ K) in the (η\eta,gg) plane. The superconducting dome is due to the shape resonance between multiple superconducting gaps where the superconducting critical temperature reaches room temperature as shown in the color plot of (TcT_{c}) as a function of the Lifshitz parameter η\eta and the electron-phonon coupling gg in the appearing Fermi surface at the ELE_{L} energy of the Lifshitz transition The critical temperature reaches room temperature superconductivity where the multi-gap superconductor is close to a Lifshitz transition for neck disrupting with the Lifshitz parameter in the range 0.6<η<10.6<\eta<1 and e-ph coupling close to 0.4.

In order to reproduce room-temperature superconductivity in CSHxCSH_{x} we have numerically evaluated the gaps and critical temperature for the case (B) where we have decreased the hopping between the wires to simulate the modified spacer material in CSHxCSH_{x} in comparison with H3SH_{3}S. Therefore we have used the transversal dispersion ΔE=145\Delta E=145 meV to simulate the superconducting dome of CSHxCSH_{x} . The results are shown in Fig.10 and Fig.11. In the case B the critical temperature TcmaxT_{c\ max} at the top of the superconductingsuperconducting domedome reaches room-temperature superconductivity.

Tuning the chemical potential, μ\mu, in the proximity of the band-edge, the superconducting system reaches different regimes which are distinguished by the Lifshitz parameter. At the Lifshitz transition for appearing of the new Fermi surface spot the Fermi level in the hot spot is very low and therefore the few electrons there are strongly coupled with lattice phonons showing the Khon anomaly and softening with superconductivity competing with charge density wave (CDW) and phase separation as it has been observed in doped diborides [Bauer et al., 2001,Agrestini et al., 2004] which show phonon softening at the maximum TcT_{c} [Simonelli et al., 2009].

In the Lifshitz transition for the topological transition of the type opening a neck the Fano-Feshbach resonance gives the maximum TcT_{c}. In fact the BCS condensate, made of the majority of electrons in the first subband, coexists with a minority of electrons in the second subband forming a condensate in BCS-BEC crossover [Guidini et al., 2016; Ochi et al., 2021]. These results show that the maximum critical temperature in the multi-gap superconducting scenario can reaches room-temperature superconductivity driven by the exchange interaction between different condensates, neglected in the BCS approximation.

V Conclusions

In conclusion, we have shown that the theory of multi-gap superconductivity in a superlattice of nanoscale stripes, which was first proposed for high temperature superconductivity in hole doped cuprate perovskites [Perali et al., 1996; Valletta et al., 1997; Bianconi et al., 1998; Bianconi, 2006; Perali et al., 2012], could provide a quantitative description of room temperature superconductivity in pressurized hydrides. We have calculated the superconducting domes for two different cases of the heterostructure of quantum stripes with larger and smaller hopping between stripes where the critical temperature is determined by both the Lifshitz parameter and the variable electron-phonon coupling in the appearing Fermi surface. The key point of our work is the solution of the Bogoljubov gap equations in a multi-gap system including the Fano-Feshbach resonance driven by the variable exchange interaction between condensates, which is usually neglected in the standard Migdal-Eliashberg theory. We have shown that multiple gaps in large Fermi surfaces with high Fermi energy in the weak coupling regime can be amplified by exchange interaction with a large gap in the strong coupling regime in a small Fermi surface spot. We have presented cases where the Fano-Feshbach resonance appears by tuning the chemical potential near an electronic topological Lifshitz transition in heterostructures of quantum wires. We have presented two different heterostructures of quantum wires where the critical temperature reaches the 200<Tc<300K200<T_{c}<300K range.

Acknowledgements.
We thank the staff of Department of Mathematics and Physics of Roma Tre University, the Computing Center of Institute of Microelectronics and Microsystems IMM of Italian National Research Council CNR and Supertripes-onlus for financial support of this research project.

References

  • Drozdov et al. (2015) A. Drozdov, M. Eremets, I. Troyan, V. Ksenofontov,  and S. I. Shylin, “Conventional superconductivity at 203 Kelvin at high pressures in the sulfur hydride system,” Nature 525, 73–76 (2015).
  • Bednorz and Müller (1988) J. G. Bednorz and K. A. Müller, “Perovskite-type oxides and the new approach to high-Tc superconductivity,” Reviews of Modern Physics 60, 585 (1988).
  • Gao et al. (1994) L. Gao, Y. Xue, F. Chen, Q. Xiong, R. Meng, D. Ramirez, C. Chu, J. Eggert,  and H. Mao, “Superconductivity up to 164 K in HgBaCaCuO (m= 1, 2, and 3) under quasi hydrostatic pressures,” Physical Review B 50, 4260 (1994).
  • Yamamoto et al. (2015) A. Yamamoto, N. Takeshita, C. Terakura,  and Y. Tokura, “High pressure effects revisited for the cuprate superconductor family with highest critical temperature,” Nature communications 6, 1–7 (2015).
  • Somayazulu et al. (2019) M. Somayazulu, M. Ahart, A. K. Mishra, Z. M. Geballe, M. Baldini, Y. Meng, V. V. Struzhkin,  and R. J. Hemley, “Evidence for superconductivity above 260 K in lanthanum superhydride at megabar pressures,” Physical Review Letters 122, 027001 (2019).
  • Drozdov et al. (2019) A. P. Drozdov, P. P. Kong, V. S. Minkov, S. P. Besedin, M. A. Kuzovnikov, S. Mozaffari, L. Balicas, F. F. Balakirev, D. E. Graf, V. B. Prakapenka, et al., “Superconductivity at 250 K in lanthanum hydride under high pressures,” Nature 569, 528–531 (2019).
  • Troyan et al. (2021) I. A. Troyan, D. V. Semenok, A. G. Kvashnin, A. V. Sadakov, O. A. Sobolevskiy, V. M. Pudalov, A. G. Ivanova, V. B. Prakapenka, E. Greenberg, A. G. Gavriliuk, et al., “Anomalous High-Temperature superconductivity in YH6,” Advanced Materials 33, 2006832 (2021).
  • Kong et al. (2021) P. Kong, V. S. Minkov, M. A. Kuzovnikov, A. P. Drozdov, S. P. Besedin, S. Mozaffari, L. Balicas, F. F. Balakirev, V. B. Prakapenka, S. Chariton, et al., “Superconductivity up to 243 K in the yttrium-hydrogen system under high pressure,” Nature Communications 12, 1–9 (2021).
  • Snider et al. (2020) E. Snider, N. Dasenbrock-Gammon, R. McBride, M. Debessai, H. Vindana, K. Vencatasamy, K. V. Lawler, A. Salamat,  and R. P. Dias, “Room-temperature superconductivity in a carbonaceous sulfur hydride,” Nature 586, 373–377 (2020).
  • Einaga et al. (2016) M. Einaga, M. Sakata, T. Ishikawa, K. Shimizu, M. I. Eremets, A. P. Drozdov, I. A. Troyan, N. Hirao,  and Y. Ohishi, “Crystal structure of the superconducting phase of sulfur hydride,” Nature physics 12, 835–838 (2016).
  • Goncharov et al. (2017) A. F. Goncharov, S. S. Lobanov, V. B. Prakapenka,  and E. Greenberg, “Stable high-pressure phases in the hs system determined by chemically reacting hydrogen and sulfur,” Physical Review B 95, 140101 (2017).
  • Duan et al. (2017) D. Duan, Y. Liu, Y. Ma, Z. Shao, B. Liu,  and T. Cui, “Structure and superconductivity of hydrides at high pressures,” National Science Review 4, 121–135 (2017).
  • Kruglov et al. (2017) I. Kruglov, R. Akashi, S. Yoshikawa, A. R. Oganov,  and M. M. D. Esfahani, “Refined phase diagram of the HS system with high-Tc superconductivity,” Physical Review B 96, 220101 (2017).
  • Purans et al. (2021) J. Purans, A. Menushenkov, S. Besedin, A. Ivanov, V. Minkov, I. Pudza, A. Kuzmin, K. Klementiev, S. Pascarelli, O. Mathon, et al., “Local electronic structure rearrangements and strong anharmonicity in YH3 under pressures up to 180 GPa,” Nature Communications 12, 1–10 (2021).
  • Eliashberg (1960) G. Eliashberg, “Interactions between electrons and lattice vibrations in a superconductor,” Sov. Phys. JETP 11, 696–702 (1960).
  • Dynes (1972) R. Dynes, “Mcmillan’s equation and the TcT_{c} of superconductors,” Solid State Communications 10, 615–618 (1972).
  • Duan et al. (2014) D. Duan, Y. Liu, F. Tian, D. Li, X. Huang, Z. Zhao, H. Yu, B. Liu, W. Tian,  and T. Cui, “Pressure-induced metallization of dense (H2S) 2H2 with high-Tc superconductivity,” Scientific reports 4, 6968 (2014).
  • Duan et al. (2015) D. Duan, X. Huang, F. Tian, D. Li, H. Yu, Y. Liu, Y. Ma, B. Liu,  and T. Cui, “Pressure-induced decomposition of solid hydrogen sulfide,” Physical Review B 91, 180502 (2015).
  • Durajski and Szczęśniak (2017) A. P. Durajski and R. Szczęśniak, “First-principles study of superconducting hydrogen sulfide at pressure up to 500 GPa,” Scientific Reports 7, 1–8 (2017).
  • Gorkov and Kresin (2018) L. P. Gorkov and V. Z. Kresin, “Colloquium: High pressure and road to room temperature superconductivity,” Reviews of Modern Physics 90, 011001 (2018).
  • Kostrzewa et al. (2020) M. Kostrzewa, K. Szczęśniak, A. Durajski,  and R. Szczęśniak, “From LaH10{LaH}_{10} to room–temperature superconductors,” Scientific Reports 10, 1–8 (2020).
  • Bianconi and Jarlborg (2015a) A. Bianconi and T. Jarlborg, “Lifshitz transitions and zero point lattice fluctuations in sulfur hydride showing near room temperature superconductivity,” Novel Superconducting Materials 1 (2015a).
  • Bianconi and Jarlborg (2015b) A. Bianconi and T. Jarlborg, “Superconductivity above the lowest earth temperature in pressurized sulfur hydride,” EPL (Europhysics Letters) 112, 37001 (2015b).
  • Jarlborg and Bianconi (2016) T. Jarlborg and A. Bianconi, “Breakdown of the migdal approximation at lifshitz transitions with giant zero-point motion in the h 3 s superconductor,” Scientific reports 6, 24816 (2016).
  • Quan and Pickett (2016) Y. Quan and W. E. Pickett, “Van hove singularities and spectral smearing in high-temperature superconducting H3S,” Physical Review B 93, 104526 (2016).
  • Souza and Marsiglio (2017) T. X. Souza and F. Marsiglio, “The possible role of van Hove singularities in the high-Tc of superconducting H3S,” International Journal of Modern Physics B 31, 1745003 (2017).
  • Capitani et al. (2017) F. Capitani, B. Langerome, J.-B. Brubach, P. Roy, A. Drozdov, M. Eremets, E. Nicol, J. Carbotte,  and T. Timusk, “Spectroscopic evidence of a new energy scale for superconductivity in H3S,” Nature physics 13, 859–863 (2017).
  • Fano (1935) U. Fano, “Sullo spettro di assorbimento dei gas nobili presso il limite dello spettro d’arco,” Il Nuovo Cimento (1924-1942) 12, 154–161 (1935).
  • Fano (1961) U. Fano, “Effects of configuration interaction on intensities and phase shifts,” Physical Review 124, 1866 (1961).
  • Lifshitz et al. (1960) I. Lifshitz et al., “Anomalies of electron characteristics of a metal in the high pressure region,” Sov. Phys. JETP 11, 1130–1135 (1960).
  • Volovik (2017) G. Volovik, “Topological Lifshitz transitions,” Low Temperature Physics 43, 47–55 (2017).
  • Volovik and Zhang (2017) G. Volovik and K. Zhang, “Lifshitz transitions, type-II Dirac and Weyl fermions, event horizon and all that,” Journal of Low Temperature Physics 189, 276–299 (2017).
  • Volovik (2018) G. E. Volovik, “Exotic Lifshitz transitions in topological materials,” Physics-Uspekhi 61, 89 (2018).
  • Kugel et al. (2008) K. Kugel, A. Rakhmanov, A. Sboychakov, N. Poccia,  and A. Bianconi, “Model for phase separation controlled by doping and the internal chemical pressure in different cuprate superconductors,” Physical Review B 78, 165124 (2008).
  • Bianconi et al. (2015) A. Bianconi, N. Poccia, A. Sboychakov, A. Rakhmanov,  and K. Kugel, “Intrinsic arrested nanoscale phase separation near a topological lifshitz transition in strongly correlated two-band metals,” Superconductor Science and Technology 28, 024005 (2015).
  • Bianconi (2003) A. Bianconi, “Ugo Fano and shape resonances,” in AIP Conference Proceedings, Vol. 652 (American Institute of Physics, 2003) pp. 13–18.
  • Vittorini-Orgeas and Bianconi (2009) A. Vittorini-Orgeas and A. Bianconi, “From Majorana theory of atomic autoionization to Feshbach resonances in high temperature superconductors,” Journal of Superconductivity and Novel Magnetism 22, 215–221 (2009).
  • Palumbo, Marcelli, and Bianconi (2016) F. Palumbo, A. Marcelli,  and A. Bianconi, “From the pion cloud of Tomonaga to the electron pairs of Schrieffer: many body wave functions from nuclear physics to condensed matter physics,” Journal of Superconductivity and Novel Magnetism 29, 3107–3111 (2016).
  • Bogoljubov, Tolmachov, and Širkov (1958) N. Bogoljubov, V. V. Tolmachov,  and D. Širkov, “A new method in the theory of superconductivity,” Fortschritte der physik 6, 605–682 (1958).
  • Heisenberg (1933) W. Heisenberg, “On the structure of atomic nuclei. iii,” Zeits. f. Phys 80, 587 (1933).
  • Majorana (1933) E. Majorana, “Ube die Kerntheorie,” Zeit. f Phys 82, 137–145 (1933).
  • Yukawa (1935) H. Yukawa, “On the interaction of elementary particles i,” Proceedings of the Physico-Mathematical Society of Japan. 3rd Series 17, 48–57 (1935).
  • Iachello (2013) F. Iachello, Interacting Bose-Fermi Systems in Nuclei, Vol. 10 (Springer Science & Business Media, 2013).
  • Feshbach and Iachello (1974) H. Feshbach and F. Iachello, “The interacting boson model,” Annals of Physics 84, 211–231 (1974).
  • Blatt and Weisskopf (1979) J. M. Blatt and V. F. Weisskopf, “Nuclear forces,” in Theoretical Nuclear Physics (Springer, 1979) pp. 119–167.
  • Perali et al. (1996) A. Perali, A. Bianconi, A. Lanzara,  and N. L. Saini, “The gap amplification at a shape resonance in a superlattice of quantum stripes: A mechanism for high TcT_{c},” Solid State Communications 100, 181–186 (1996).
  • Valletta et al. (1997) A. Valletta, A. Bianconi, A. Perali,  and N. Saini, “Electronic and superconducting properties of a superlattice of quantum stripes at the atomic limit,” Zeitschrift für Physik B Condensed Matter 104, 707–713 (1997).
  • Bianconi et al. (1998) A. Bianconi, A. Valletta, A. Perali,  and N. L. Saini, “Superconductivity of a striped phase at the atomic limit,” Physica C: Superconductivity 296, 269–280 (1998).
  • Bianconi (2006) A. Bianconi, “Multiband superconductivity in high TcT_{c} cuprates and diborides,” Journal of Physics and Chemistry of Solids 67, 567–570 (2006).
  • Perali et al. (2012) A. Perali, D. Innocenti, A. Valletta,  and A. Bianconi, “Anomalous isotope effect near a 2.5 Lifshitz transition in a multi-band multi-condensate superconductor made of a superlattice of stripes,” Superconductor Science and Technology 25, 124002 (2012).
  • Bianconi et al. (2001a) A. Bianconi, S. Agrestini, G. Bianconi, D. Di Castro,  and N. Saini, “A quantum phase transition driven by the electron lattice interaction gives high TcT_{c} superconductivity,” Journal of alloys and compounds 317, 537–541 (2001a).
  • Perali et al. (2004) A. Perali, P. Pieri, L. Pisani,  and G. Strinati, “BCS-BEC crossover at finite temperature for superfluid trapped fermi atoms,” Physical Review Letters 92, 220404 (2004).
  • Perali, Pieri, and Strinati (2004) A. Perali, P. Pieri,  and G. Strinati, “Quantitative comparison between theoretical predictions and experimental results for the BCS-BEC crossover,” Physical Review Letters 93, 100404 (2004).
  • Bianconi (2005) A. Bianconi, “Feshbach shape resonance in multiband superconductivity in heterostructures,” Journal of Superconductivity 18, 625–636 (2005).
  • Innocenti et al. (2010a) D. Innocenti, N. Poccia, A. Ricci, A. Valletta, S. Caprara, A. Perali,  and A. Bianconi, “Resonant and crossover phenomena in a multiband superconductor: Tuning the chemical potential near a band edge,” Physical Review B 82, 184528 (2010a).
  • Innocenti et al. (2010b) D. Innocenti, S. Caprara, N. Poccia, A. Ricci, A. Valletta,  and A. Bianconi, “Shape resonance for the anisotropic superconducting gaps near a lifshitz transition: the effect of electron hopping between layers,” Superconductor Science and Technology 24, 015012 (2010b).
  • Bianconi (2012) G. Bianconi, “Superconductor-insulator transition on annealed complex networks,” Physical Review E 85, 061113 (2012).
  • Innocenti and Bianconi (2013) D. Innocenti and A. Bianconi, “Isotope effect at the Fano resonance in superconducting gaps for multi-band superconductors at a 2.5 Lifshitz transition,” Journal of Superconductivity and Novel Magnetism 26, 1319–1324 (2013).
  • Bianconi (2013a) A. Bianconi, “Shape resonances in superstripes,” Nature Physics 9, 536–537 (2013a).
  • Mazziotti et al. (2017) M. V. Mazziotti, A. Valletta, G. Campi, D. Innocenti, A. Perali,  and A. Bianconi, “Possible Fano resonance for high-TcT_{c} multi-gap superconductivity in p-Terphenyl doped by K at the Lifshitz transition,” EPL (Europhysics Letters) 118, 37003 (2017).
  • Pinto et al. (2020) N. Pinto, C. Di Nicola, A. Trapananti, M. Minicucci, A. Di Cicco, A. Marcelli, A. Bianconi, F. Marchetti, C. Pettinari,  and A. Perali, “Potassium-doped para-terphenyl: Structure, electrical transport properties and possible signatures of a superconducting transition,” Condensed Matter 5, 78 (2020).
  • Mazziotti et al. (2021a) M. V. Mazziotti, A. Valletta, R. Raimondi,  and A. Bianconi, “Multigap superconductivity at an unconventional Lifshitz transition in a three-dimensional Rashba heterostructure at the atomic limit,” Phys. Rev. B 103, 024523 (2021a).
  • Bianconi et al. (2014) A. Bianconi, D. Innocenti, A. Valletta,  and A. Perali, “Shape resonances in superconducting gaps in a 2DEG at oxide-oxide interface,” in J. Phys.: Conf. Ser., Vol. 529 (IOP Publishing, 2014) p. 012007.
  • Mazziotti et al. (2018) M. V. Mazziotti, N. Scopigno, M. Grilli,  and S. Caprara, “Majorana Fermions in One-Dimensional Structures at LaAlO3/SrTiO3 Oxide Interfaces,” Condensed Matter 3, 37 (2018).
  • Feshbach (1962) H. Feshbach, “A unified theory of nuclear reactions. ii,” Annals of Physics 19, 287–313 (1962).
  • Feshbach (1983) H. Feshbach, “Conference comments,” Nuclear Physics A 409, 423–428 (1983).
  • Feshbach (2014) H. Feshbach, “Nuclear physics: Comments and reflections,” Niels Bohr: Physics and the World , 117 (2014).
  • Miroshnichenko, Flach, and Kivshar (2010) A. E. Miroshnichenko, S. Flach,  and Y. S. Kivshar, “Fano resonances in nanoscale structures,” Reviews of Modern Physics 82, 2257 (2010).
  • Bianconi et al. (1978) A. Bianconi, H. Petersen, F. C. Brown,  and R. Bachrach, “K-shell photoabsorption spectra of N2 and N2O using synchrotron radiation,” Physical Review A 17, 1907 (1978).
  • Bianconi (1980) A. Bianconi, “Surface X-ray absorption spectroscopy: Surface EXAFS and surface XANES,” Applications of Surface Science 6, 392–418 (1980).
  • Bianconi et al. (1982) A. Bianconi, M. Dell’Ariccia, P. Durham,  and J. Pendry, “Multiple-scattering resonances and structural effects in the x-ray-absorption near-edge spectra of Fe II and Fe III hexacyanide complexes,” Physical Review B 26, 6502 (1982).
  • Suhl, Matthias, and Walker (1959) H. Suhl, B. Matthias,  and L. Walker, “Bardeen-cooper-schrieffer theory of superconductivity in the case of overlapping bands,” Physical Review Letters 3, 552 (1959).
  • Moskalenko (1959) V. Moskalenko, “Superconductivity in metals with overlapping energy bands,” Fiz. Metal. Metalloved 8, 2518–2520 (1959).
  • Kondo (1963) J. Kondo, “Superconductivity in transition metals,” Progress of Theoretical Physics 29, 1–9 (1963).
  • Mazziotti et al. (2021b) M. V. Mazziotti, T. Jarlborg, A. Bianconi,  and A. Valletta, “Room temperature superconductivity dome at a Fano resonance in superlattices of wires,” EPL (Europhysics Letters) 134, 17001 (2021b).
  • Zwierlein et al. (2004) M. Zwierlein, C. Stan, C. Schunck, S. Raupach, A. Kerman,  and W. Ketterle, “Condensation of pairs of fermionic atoms near a feshbach resonance,” Physical Review Letters 92, 120403 (2004).
  • Zhang et al. (2004) J. Zhang, E. Van Kempen, T. Bourdel, L. Khaykovich, J. Cubizolles, F. Chevy, M. Teichmann, L. Tarruell, S. Kokkelmans,  and C. Salomon, “P-wave Feshbach resonances of ultracold Li6{Li}_{6},” Physical Review A 70, 030702 (2004).
  • Uemura et al. (1989) Y. Uemura, G. Luke, B. Sternlieb, J. Brewer, J. Carolan, W. Hardy, R. Kadono, J. Kempton, R. Kiefl, S. Kreitzman, et al., “Universal correlations between TcT_{c} and n s m*(carrier density over effective mass) in high-TcT_{c} cuprate superconductors,” Physical Review Letters 62, 2317 (1989).
  • Pistolesi and Strinati Calvanese (1994) F. Pistolesi and G. Strinati Calvanese, “Evolution from BCS superconductivity to Bose condensation: role of the parameter kFk_{F}ξ\xi,” Physical Review B 49, 6356 (1994).
  • Agrestini et al. (2003) S. Agrestini, N. Saini, G. Bianconi,  and A. Bianconi, “The strain of CuO2 lattice: the second variable for the phase diagram of cuprate perovskites,” Journal of Physics A: Mathematical and General 36, 9133 (2003).
  • Szczesniak and Durajski (2017) R. Szczesniak and A. Durajski, “The isotope effect in H3S superconductor,” Solid State Communications 249, 30–33 (2017).
  • Szczesniak and Durajski (2018) R. Szczesniak and A. P. Durajski, “Unusual sulfur isotope effect and extremely high critical temperature in H3S superconductor,” Scientific reports 8, 1–9 (2018).
  • Perali et al. (1997) A. Perali, A. Valletta, G. Bardeiloni, A. Bianconi, A. Lanzara,  and N. Saini, “The isotope effect in a superlattice of quantum stripes,” Journal of Superconductivity 10, 355–359 (1997).
  • Bianconi, Innocenti, and Campi (2013) A. Bianconi, D. Innocenti,  and G. Campi, “Superstripes and superconductivity in complex granular matter,” Journal of Superconductivity and Novel Magnetism 26, 2585–2588 (2013).
  • Bianconi (2013b) A. Bianconi, “Shape resonances in multi-condensate granular superconductors formed by networks of nanoscale-striped puddles,” in J. Phys.: Conf. Ser., Vol. 449 (IOP Publishing, 2013) p. 012002.
  • Masella et al. (2019) G. Masella, A. Angelone, F. Mezzacapo, G. Pupillo,  and N. V. Prokof?ev, “Supersolid stripe crystal from finite-range interactions on a lattice,” Physical Review Letters 123, 045301 (2019).
  • Testardi (1975) L. Testardi, “Structural instability and superconductivity in A-15 compounds,” Reviews of Modern Physics 47, 637 (1975).
  • Bianconi et al. (1991) A. Bianconi, S. Della Longa, C. Li, M. Pompa, A. Congiu-Castellano, D. Udron, A. Flank,  and P. Lagarde, “Linearly polarized Cu L3L_{3}-edge x-ray-absorption near-edge structure of Bi2CaSr2Cu2O8,” Physical Review B 44, 10126 (1991).
  • Bianconi et al. (1996a) A. Bianconi, N. Saini, A. Lanzara, M. Missori, T. Rossetti, H. Oyanagi, H. Yamaguchi, K. Oka,  and T. Ito, “Determination of the Local Lattice Distortions in the CuO2 Plane of La1.85Sr0.15CuO4,” Physical Review Letters 76, 3412 (1996a).
  • Bianconi et al. (1996b) A. Bianconi, M. Lusignoli, N. Saini, P. Bordet, Å. Kvick,  and P. Radaelli, “Stripe structure of the CuO2 plane in Bi2Sr2CaCu2O8+y by anomalous X-ray diffraction,” Physical Review B 54, 4310 (1996b).
  • Gavrichkov et al. (2019) V. A. Gavrichkov, Y. Shanḱo, N. G. Zamkova,  and A. Bianconi, “Is there any hidden symmetry in the stripe structure of perovskite high-temperature superconductors?” The Journal of Physical Chemistry Letters 10, 1840–1844 (2019).
  • Bianconi et al. (2000) A. Bianconi, N. L. Saini, S. Agrestini, D. Di Castro,  and G. Bianconi, “The strain quantum critical point for superstripes in the phase diagram of all cuprate perovskites,” International Journal of Modern Physics B 14, 3342–3355 (2000).
  • Di Castro et al. (2000) D. Di Castro, G. Bianconi, M. Colapietro, A. Pifferi, N. Saini, S. Agrestini,  and A. Bianconi, “Evidence for the strain critical point in high TcT_{c} superconductors,” The European Physical Journal B-Condensed Matter and Complex Systems 18, 617–624 (2000).
  • Barba et al. (2018) L. Barba, G. Chita, G. Campi, L. Suber, E. M. Bauer, A. Marcelli,  and A. Bianconi, “Anisotropic thermal expansion of p-terphenyl: A self-assembled supramolecular array of poly-p-phenyl nanoribbons,” Journal of Superconductivity and Novel Magnetism 31, 703–710 (2018).
  • Bauer et al. (2001) E. Bauer, C. Paul, S. Berger, S. Majumdar, H. Michor, M. Giovannini, A. Saccone,  and A. Bianconi, “Thermal conductivity of superconducting MgB2,” Journal of Physics: Condensed Matter 13, L487 (2001).
  • Agrestini et al. (2001) S. Agrestini, D. Di Castro, M. Sansone, N. Saini, A. Saccone, S. De Negri, M. Giovannini, M. Colapietro,  and A. Bianconi, “High TcT_{c} superconductivity in a critical range of micro-strain and charge density in diborides,” Journal of Physics: Condensed Matter 13, 11689 (2001).
  • Bianconi et al. (2001b) A. Bianconi, D. Di Castro, S. Agrestini, G. Campi, N. Saini, A. Saccone, S. De Negri,  and M. Giovannini, “A superconductor made by a metal heterostructure at the atomic limit tuned at the shape resonance’: MgB2MgB_{2},” Journal of Physics: Condensed Matter 13, 7383 (2001b).
  • Agrestini et al. (2004) S. Agrestini, C. Metallo, M. Filippi, L. Simonelli, G. Campi, C. Sanipoli, E. Liarokapis, S. De Negri, M. Giovannini, A. Saccone, et al., “Substitution of Sc for Mg in MgB2: Effects on transition temperature and kohn anomaly,” Physical Review B 70, 134514 (2004).
  • Di Castro et al. (2002) D. Di Castro, S. Agrestini, G. Campi, A. Cassetta, M. Colapietro, A. Congeduti, A. Continenza, S. De Negri, M. Giovannini, S. Massidda, et al., “The amplification of the superconducting TcT_{c} by combined effect of tuning of the Fermi level and the tensile micro-strain in Al1-x Mgx B2,” EPL (Europhysics Letters) 58, 278 (2002).
  • Ricci et al. (2009) A. Ricci, N. Poccia, G. Ciasca, M. Fratini,  and A. Bianconi, “The microstrain-doping phase diagram of the iron pnictides: heterostructures at atomic limit,” Journal of Superconductivity and Novel Magnetism 22, 589–593 (2009).
  • Ricci et al. (2010) A. Ricci, N. Poccia, B. Joseph, L. Barba, G. Arrighetti, G. Ciasca, J.-Q. Yan, R. W. McCallum, T. A. Lograsso, N. Zhigadlo, et al., “Structural phase transition and superlattice misfit strain of RFeAsO (R= La, Pr, Nd, Sm),” Physical Review B 82, 144507 (2010).
  • Kordyuk (2018) A. A. Kordyuk, “Electronic band structure of optimal superconductors: from cuprates to ferropnictides and back again,” Low Temperature Physics 44, 477–486 (2018).
  • Pustovit and Kordyuk (2016) Y. V. Pustovit and A. Kordyuk, “Metamorphoses of electronic structure of FeSe-based superconductors,” Low Temperature Physics 42, 995–1007 (2016).
  • Guidini et al. (2016) A. Guidini, L. Flammia, M. V. Milošević,  and A. Perali, “BCS-BEC crossover in quantum confined superconductors,” Journal of Superconductivity and Novel Magnetism 29, 711–715 (2016).
  • Cariglia et al. (2016) M. Cariglia, A. Vargas-Paredes, M. M. Doria, A. Bianconi, M. V. Milošević,  and A. Perali, “Shape-resonant superconductivity in nanofilms: from weak to strong coupling,” Journal of Superconductivity and Novel Magnetism 29, 3081–3086 (2016).
  • Doria, Cariglia, and Perali (2016) M. M. Doria, M. Cariglia,  and A. Perali, “Multigap superconductivity and barrier-driven resonances in superconducting nanofilms with an inner potential barrier,” Physical Review B 94, 224513 (2016).
  • Bussmann-Holder et al. (2016) A. Bussmann-Holder, J. Köhler, M.-H. Whangbo, A. Bianconi,  and A. Simon, “High temperature superconductivity in sulfur hydride under ultrahigh pressure: A complex superconducting phase beyond conventional BCS,” Novel Superconducting Materials 1 (2016).
  • Valentinis, Van Der Marel, and Berthod (2016) D. Valentinis, D. Van Der Marel,  and C. Berthod, “BCS superconductivity near the band edge: Exact results for one and several bands,” Physical Review B 94, 024511 (2016).
  • Bussmann-Holder et al. (2017) A. Bussmann-Holder, J. Kohler, A. Simon, M.-H. Whangbo, A. Bianconi,  and A. Perali, “The road map toward room-temperature superconductivity: Manipulating different pairing channels in systems composed of multiple electronic components,” Condensed Matter 2, 24 (2017).
  • Bussmann-Holder et al. (2019) A. Bussmann-Holder, H. Keller, A. Simon,  and A. Bianconi, “Multi-band superconductivity and the steep band/flat band scenario,” Condensed Matter 4, 91 (2019).
  • Chubukov and Mozyrsky (2018) A. V. Chubukov and D. Mozyrsky, “Evolution of the dynamics of neutral superconductors between BCS and BEC regimes: The variational approach,” Low Temperature Physics 44, 528–533 (2018).
  • Salasnich et al. (2019) L. Salasnich, A. Shanenko, A. Vagov, J. A. Aguiar,  and A. Perali, “Screening of pair fluctuations in superconductors with coupled shallow and deep bands: A route to higher-temperature superconductivity,” Physical Review B 100, 064510 (2019).
  • Kagan and Bianconi (2019) M. Y. Kagan and A. Bianconi, “Fermi-Bose mixtures and BCS-BEC crossover in high-Tc superconductors,” Condensed Matter 4, 51 (2019).
  • Tajima, Perali, and Pieri (2020) H. Tajima, A. Perali,  and P. Pieri, “BCS-BEC Crossover and Pairing Fluctuations in a Two Band Superfluid/Superconductor: A T Matrix Approach,” Condensed Matter 5(1), 10 (2020).
  • Vargas-Paredes et al. (2020) A. Vargas-Paredes, A. Shanenko, A. Vagov, M. Milošević,  and A. Perali, “Crossband versus intraband pairing in superconductors: Signatures and consequences of the interplay,” Physical Review B 101, 094516 (2020).
  • Bianconi (1996) A. Bianconi, “High TcT_{c} superconductors made by metal heterostuctures at the atomic limit,” European Patent EP0733271 Sept 25 (1996).
  • Bianconi (2001) A. Bianconi, “Process of increasing the critical temperature TcT_{c} of a bulk superconductor by making metal heterostructures at the atomic limit,”  (2001), US Patent 6,265,019.
  • Bianconi (1994) A. Bianconi, “On the possibility of new high-Tc superconductors by producing metal heterostructures as in the cuprate perovskites,” Solid State Communications 89, 933–936 (1994).
  • Isobe and Fu (2019) H. Isobe and L. Fu, “Supermetal,” Physical Review Research 1, 033206 (2019).
  • Bianconi et al. (1997) A. Bianconi, A. Valletta, A. Perali,  and N. Saini, “High TcT_{c} superconductivity in a superlattice of quantum stripes,” Solid State Communications 102, 369–374 (1997).
  • Migdal (1958) A. Migdal, “Interaction between electrons and lattice vibrations in a normal metal,” Sov. Phys. JETP 7, 996–1001 (1958).
  • Simonelli et al. (2009) L. Simonelli, V. Palmisano, M. Fratini, M. Filippi, P. Parisiades, D. Lampakis, E. Liarokapis,  and A. Bianconi, “Isotope effect on the E2g phonon and mesoscopic phase separation near the electronic topological transition in Mg1-xAlxB2,” Physical Review B 80, 014520 (2009).
  • Ochi et al. (2021) K. Ochi, H. Tajima, K. Iida,  and H. Aoki, “Resonant pair-exchange scattering and BCS-BEC crossover in a system composed of dispersive and heavy incipient bands: a Feshbach analogy,” arXiv preprint arXiv:2107.13805  (2021).