Preprint
Concept Paper

This version is not peer-reviewed.

On the Theoretical Inconsistencies of Cooper Pairing in Superconductivity

Submitted:

03 December 2025

Posted:

05 December 2025

You are already at the latest version

Abstract
The BCS theory of superconductivity, which relies on the formation of Cooper pairs mediated by lattice phonons, has stood for decades as the cornerstone of our understanding of superconductivity in conventional metals. However, critical inspection reveals that several theoretical and experimental inconsistencies persist in this framework, especially when extended to high-temperature and unconventional superconductors. This paper rigorously analyzes these inconsistencies, with emphasis on the inadequacy of phonon-mediated interactions to overcome Coulomb repulsion, the questionable nature of the long-range coherence implied by the size of Cooper pairs, and the breakdown of BCS predictions in strongly correlated systems. We present a calculation-intensive critique, highlighting the need for a deeper, possibly non-phononic mechanism for electron pairing or collective quantum behavior in superconductors. The BCS theory of superconductivity, premised on the formation of Cooper pairs via weak electron-phonon coupling, has long served as the canonical framework for understanding low-temperature superconductors. However, we argue that this framework is conceptually and physically insufficient—even for conventional materials. This paper presents a detailed theoretical critique grounded in explicit calculations, exposing contradictions in the length and energy scales involved, the lack of real-space localization of paired electrons, and the incompatibility between the BCS ground state and a physically bound pair in position space. We emphasize that the superconducting energy gap may better reflect a many-body correlation scale rather than a two-body binding energy. Further, we discuss topological and quantum field theoretic obstructions to pairing and reframe superconductivity as a macroscopic quantum coherent state independent of pair formation. Our approach challenges the narrative that Cooper pairing is a necessary cause of superconductivity and instead highlights the role of collective phase coherence, entanglement, and broken gauge symmetry as possible fundamental mechanisms.
Keywords: 
;  ;  ;  

1. Introduction

The Bardeen-Cooper-Schrieffer (BCS) theory of superconductivity has achieved remarkable success in explaining several properties of low-temperature superconductors, including the energy gap, specific heat anomalies, and the Meissner effect. However, the foundational mechanism — that of electron pairing via phonon-mediated attraction — rests upon assumptions that begin to falter under rigorous scrutiny and when applied to unconventional materials. The concept of the Cooper pair, introduced by Cooper in 1956 [1], assumes that two electrons with opposite spin and momentum can form a bound state via an effective attractive potential mediated by lattice vibrations, even in the presence of a strong Coulomb repulsion.
Let us begin by stating the canonical expression for the BCS energy gap Δ , derived via mean-field approximation:
Δ ( T ) = Δ ( 0 ) tanh π k B T c Δ ( 0 ) T c T 1
This expression fits experimental data for low-temperature superconductors but becomes inadequate for high- T c materials. Furthermore, the underlying assumptions of weak coupling and long coherence lengths are not upheld in many real systems, particularly in cuprate and iron-based superconductors.
The Bardeen-Cooper-Schrieffer (BCS) theory of superconductivity has long been regarded as a milestone in condensed matter physics, offering elegant solutions to low-temperature superconducting phenomena such as the energy gap, the Meissner effect, and the temperature dependence of specific heat. Central to this framework is the formation of Cooper pairs—composite bosons resulting from phonon-mediated attractive interactions between electrons near the Fermi surface. Despite its widespread acceptance, critical analysis reveals foundational and interpretational issues that cast doubt on the universal applicability of this formalism.
The canonical BCS expression for the superconducting energy gap Δ ( T ) is typically given by:
Δ ( T ) = Δ ( 0 ) tanh π k B T c Δ ( 0 ) T c T 1
where T c is the critical temperature and k B is Boltzmann’s constant. While this expression shows reasonable agreement with experimental data for conventional superconductors such as Al, Sn, or Pb, its derivation is contingent upon weak-coupling approximations and assumes a uniform Fermi liquid background, which may not be accurate for all metallic systems.
Moreover, the characteristic coherence length ξ 0 derived from BCS theory:
ξ 0 = v F π Δ
yields values on the order of tens to hundreds of nanometers in low-temperature superconductors. This implies that the Cooper pair wavefunction extends over thousands of lattice sites. Such delocalization is difficult to reconcile with any real-space notion of a “bound pair,” and introduces ambiguity in interpreting the pair as a localized molecular entity.
Another challenge arises when comparing the small superconducting gap, typically Δ 1 meV , to the Fermi energy, E F 1 10 eV , leading to an extreme disparity:
Δ E F 10 3 10 4
Such a small perturbation energy raises questions regarding the physical robustness of the superconducting state, especially in the presence of environmental fluctuations, weak magnetic fields, or moderate current densities. If the superconducting state is truly energetically favored, it remains puzzling why it collapses under perturbations far smaller than the Fermi energy scale.
Additionally, the BCS ground state is a momentum-space construction:
| Ψ BCS = k u k + v k c k c k | 0
This entangled state does not correspond to real-space localized electron pairs, but rather a collective condensate over the Fermi sea. Consequently, the commonly used metaphor of “pairing” should not be confused with a literal two-body wavefunction, but instead as a many-body correlated state. Yet, most pedagogical treatments and popular descriptions continue to emphasize the real-space pairing picture, which could be misleading even in low-temperature superconductors.
In light of these inconsistencies, this paper revisits the BCS formalism with a critical lens. We examine the mismatch between drift velocity and signal propagation, the fragility of the superconducting phase, the absence of direct detection of Cooper pairs, and the possible alternative descriptions grounded in symmetry breaking, macroscopic entanglement, or topological constraints. Our goal is not merely to question the numerical validity of BCS predictions, but to critically assess its interpretational and foundational robustness.

2. Energy Scale Limitations and Coulomb Repulsion

A major issue with the Cooper pair concept lies in the inconsistency between the energy scales of phonons and Coulomb interaction. Phonon-mediated interaction energy is typically of the order of tens of meV. In contrast, Coulomb repulsion between two electrons separated by a few angstroms is much larger. For instance, the Coulomb potential between two electrons at a distance of a = 2 Å is:
U C = e 2 4 π ϵ 0 a ( 1.6 × 10 19 ) 2 4 π ( 8.85 × 10 12 ) × 2 × 10 10 7.2 eV
On the other hand, the Debye energy ω D in most metals is:
ω D 30 50 meV
This presents an energy scale mismatch by over two orders of magnitude. In BCS theory, the effective attraction is assumed to overcome this repulsion. The effective potential V leads to a bound state only when the density of states N ( 0 ) and V satisfy:
1 = V N ( 0 ) ln 2 ω D Δ
However, unless V is unphysically large or N ( 0 ) is anomalously high, this equation leads to exponentially small values of Δ , which is incompatible with observed high-temperature gaps.

3. Coherence Length and Long-Range Entanglement

The coherence length in a BCS superconductor is given by:
ξ = v F π Δ
For typical metals, taking v F 10 6 m / s and Δ 1 meV :
ξ ( 1.05 × 10 34 ) × ( 10 6 ) π × ( 1.6 × 10 22 ) 2.1 × 10 7 m = 210 nm
This implies that Cooper pairs are extended over hundreds of nanometers. The physical feasibility of two electrons being coherently entangled over such a distance, especially in the presence of impurities, thermal noise, and lattice disorder, remains questionable. Furthermore, it violates the intuitive notion of localized interactions mediated by short-wavelength phonons.

4. Failure of BCS in High- T c Superconductors

High-temperature superconductors, particularly the cuprates, display behaviors that are irreconcilable with BCS theory. For instance, the isotope effect, a cornerstone of phonon-based pairing, is absent or reversed in several high- T c materials [2]. Moreover, the pairing symmetry in these systems is d-wave, rather than the s-wave predicted by conventional BCS theory.
Angle-resolved photoemission spectroscopy (ARPES) and scanning tunneling microscopy (STM) reveal pseudogaps and anomalous normal states in these materials, suggesting a fundamentally different mechanism. In particular, the observed gap magnitude Δ in YBa2Cu3O7−δ is around 20–30 meV, inconsistent with predictions from Eq. (4) unless one assumes a non-phononic, strong coupling mechanism.
Furthermore, the phase diagram of these materials shows that superconductivity arises from a strange metallic state, not a Fermi liquid. Therefore, any theory based on a weakly interacting Fermi liquid is inadequate for capturing the full phenomenology of high- T c superconductivity.

5. Alternative Theoretical Considerations

Given the challenges faced by BCS theory, numerous alternatives have been proposed. Anderson’s resonating valence bond (RVB) theory [3] suggests that superconductivity emerges from a spin-liquid ground state, where electron pairing is due to quantum spin correlations rather than phonons.
Topological superconductivity and holographic dualities from AdS/CFT have also been proposed, especially for materials with nontrivial band structures and strong spin-orbit coupling. These models predict gapless edge modes and Majorana states that are not accommodated within BCS.
Another alternative involves bipolaron formation, where strong electron-lattice interactions trap electrons in pairs, but these mechanisms suffer from mobility issues. Nonetheless, the very existence of viable alternatives underlines the theoretical non-universality of the Cooper pair picture.

6. Temperature Scale Paradox

The central argument of BCS theory relies on the idea that a weak phonon-mediated attractive interaction can bind two electrons into a Cooper pair despite their natural electrostatic repulsion. However, this assumption demands scrutiny when examining the energy scales involved in the problem. The paradox arises from a stark disparity between the magnitudes of phonon energies and Coulomb interaction energies. Phonons, being quantized lattice vibrations, typically operate in the range of 10–100 meV, while Coulomb repulsion operates in the range of several electronvolts. This significant mismatch in energy scales raises doubts about the plausibility of the phonon-mediated pairing mechanism, especially in strongly correlated and high-temperature superconductors.
To analyze this discrepancy quantitatively, consider the Coulomb repulsion between two electrons separated by a typical lattice spacing of a = 2 Å = 2 × 10 10 m . The electrostatic potential energy between them is given by:
U C = e 2 4 π ϵ 0 a
Substituting known constants, e = 1.6 × 10 19 C and ϵ 0 = 8.85 × 10 12 F / m , we obtain:
U C = ( 1.6 × 10 19 ) 2 4 π × ( 8.85 × 10 12 ) × ( 2 × 10 10 ) 7.2 eV
On the other hand, the typical Debye frequency ω D in metals like lead, aluminum, and niobium is on the order of 10 13 Hz . The associated phonon energy is:
E D = ω D
Using = 1.05 × 10 34 J · s and ω D = 5 × 10 13 Hz :
E D = 1.05 × 10 34 × 5 × 10 13 = 5.25 × 10 21 J 32.8 meV
Thus, the phonon energy is nearly two orders of magnitude smaller than the Coulomb repulsion. This raises the question of how a phonon could mediate sufficient attraction to overcome such a repulsive force. The energy imbalance implies that a purely phonon-based mechanism should be incapable of forming a bound state under typical metallic conditions.
BCS theory circumvents this objection by invoking the retardation effect. The phonon interaction is delayed, whereas the Coulomb interaction is instantaneous. Hence, the effective interaction between two electrons near the Fermi surface can be attractive due to the time lag. Mathematically, this is introduced via the Eliashberg function α 2 F ( ω ) in the extended Migdal-Eliashberg theory, where the effective coupling constant λ is defined as:
λ = 2 0 α 2 F ( ω ) ω d ω
However, this formulation assumes that Migdal’s theorem holds, which requires that the ratio:
ω D E F 1
where E F is the Fermi energy. In high- T c superconductors and other strongly correlated systems, this condition breaks down. For instance, in optimally doped YBa2Cu3O7−δ, E F is estimated to be only a few hundred meV, and ω D can be a significant fraction of this, invalidating the approximation.
Moreover, experimental studies [4,5] have shown that even in conventional superconductors, the Coulomb pseudopotential μ * must be fine-tuned to produce realistic critical temperatures. The McMillan formula:
T c = θ D 1.45 exp 1.04 ( 1 + λ ) λ μ * ( 1 + 0.62 λ )
demonstrates this sensitivity. Here, μ * represents the screened Coulomb repulsion. Typical values used in modeling range from 0.1 to 0.15. If μ * is underestimated, the calculated T c becomes negligible, reaffirming that phonon-mediated attraction alone is insufficient without fine-tuned cancellation of Coulomb effects.
Furthermore, numerical simulations within density functional theory (DFT) have highlighted that the phonon spectrum does not soften sufficiently near the superconducting transition temperature T c to account for strong enough electron-lattice interactions. Inelastic neutron scattering studies [6] in materials like niobium and lead have failed to show dramatic phonon anomalies near T c , undermining the assertion that phonons play a dominant role in pairing.
Lastly, from a thermodynamic standpoint, the free energy reduction associated with the superconducting transition must balance the condensation energy of Cooper pairs. The energy gain per Cooper pair is:
Δ E = 1 2 N ( 0 ) Δ 2
Using Δ 1.5 meV for lead and N ( 0 ) 10 47 J 1 m 3 , the condensation energy density is:
U cond = 1 2 × 10 47 × ( 2.4 × 10 22 ) 2 2.9 × 10 3 J / m 3
This value is small compared to the total electronic energy density, further suggesting that the energy balance required to form Cooper pairs via weak phonon interactions is marginal.
In summary, the phonon-based mechanism for electron pairing suffers from a severe energy scale disparity when compared with Coulomb repulsion. Theoretical workarounds such as retardation effects and screened Coulomb potentials require delicate fine-tuning and lose validity in strongly correlated materials. Therefore, the assumption that weak lattice vibrations can universally lead to electron pairing appears physically implausible under many realistic conditions.

7. Length Scale Discrepancy

A fundamental inconsistency in the BCS framework arises when comparing the length scale over which electron pairing is said to occur with the intrinsic range of the phonon-mediated interaction. The coherence length ξ , which characterizes the spatial extent of a Cooper pair, is typically on the order of 100 nm for conventional superconductors like aluminum and lead. However, the interaction responsible for pairing, i.e., the electron-phonon coupling, is inherently short-range and local.
The BCS coherence length is given by:
ξ 0 = v F π Δ
Using v F 1 × 10 6 m / s and Δ 1.5 meV = 2.4 × 10 22 J for lead, and = 1.05 × 10 34 J · s , we compute:
ξ 0 = 1.05 × 10 34 × 10 6 π × 2.4 × 10 22 1.39 × 10 7 m = 139 nm
This coherence length spans hundreds of lattice constants. In comparison, the spatial range of phonon-mediated interaction is typically on the order of the lattice constant, a 2 4 Å .
This presents a significant discrepancy, as it is unclear how such a short-range interaction can mediate coherent binding over tens to hundreds of nanometers. The typical phonon wavelength λ at the Debye frequency ω D is on the order of:
λ D = 2 π v s ω D
Assuming v s 3 × 10 3 m / s and ω D 5 × 10 13 Hz , we find:
λ D 2 π × 3 × 10 3 5 × 10 13 3.77 × 10 10 m = 3.77 Å
The phonon interaction is thus highly localized in real space, acting over interatomic distances, while the Cooper pair wavefunction extends over hundreds of nanometers. This is further complicated by the fact that the interaction is mediated through the lattice and is thus sensitive to impurities and defects.
In a disordered system, the mean free path l of electrons is reduced due to scattering. When l < ξ , coherence is expected to break down, yet superconductivity persists even in relatively dirty samples. This paradox is addressed in Anderson’s theorem [7], which argues that time-reversal symmetry, rather than translational symmetry, is essential for preserving superconductivity in the presence of nonmagnetic disorder. However, this theorem assumes the pairing potential remains homogeneous.
The spatial decay of the pair correlation function F ( r ) in real space is often modeled as:
F ( r ) sin ( k F r ) k F r e r / ξ
Even though the exponential decay length ξ is long, the oscillatory factor sin ( k F r ) induces rapid phase changes. For k F 1.2 × 10 10 m 1 , the oscillation period is:
λ F = 2 π k F 5.2 Å
Therefore, within a single coherence length, the pair wavefunction undergoes 2000 oscillations. Maintaining phase coherence over such oscillations is physically challenging, especially when the medium is not perfectly crystalline. This raises questions about the robustness of long-range coherence and the feasibility of constructing a pair wavefunction that maintains its quantum coherence in such environments.
Moreover, tunneling experiments such as those by Giaever [8] demonstrate that the gap structure is sharp only in very clean samples. In disordered systems, the broadening of the gap structure suggests loss of coherence. The fact that BCS theory still applies under such conditions is not fully explained without invoking approximations or idealized models.
Furthermore, microscopic simulations based on real-space Bogoliubov-de Gennes equations [9] show that the order parameter varies significantly on the scale of the coherence length in inhomogeneous systems. This spatial fluctuation further complicates the notion of uniform pairing over long distances via a short-range interaction.
In summary, the physical mechanism by which short-range phonon interactions generate long-range coherent pairs is not satisfactorily explained within the BCS framework. The large discrepancy between the coherence length and the range of interaction, combined with the susceptibility to decoherence in real materials, suggests that the current understanding of Cooper pairing may be incomplete or oversimplified.

8. Violation of Electron Identity (Fermionic Nature)

One of the most profound conceptual transitions in BCS theory is the assumption that two fermions — electrons — can form a composite boson, known as a Cooper pair. This pair is then treated as a bosonic entity capable of condensing into a macroscopic quantum state. While this transformation is mathematically permissible under specific constraints, it creates interpretational challenges when examined from the standpoint of fundamental particle identity and thermal stability.
Fermions obey the Pauli exclusion principle and Fermi-Dirac statistics. That is, for two fermions ψ 1 and ψ 2 , the antisymmetric wavefunction under exchange is:
Ψ ( r 1 , r 2 ) = Ψ ( r 2 , r 1 )
However, a pair of spin- 1 / 2 fermions can form a composite boson if their total spin is an integer, and if their combined wavefunction is symmetric. This is the justification behind considering Cooper pairs as bosons. The BCS wavefunction is expressed as:
| Ψ BCS = k ( u k + v k c k c k ) | 0
This wavefunction allows for a coherent superposition of pair states over momentum space. The amplitude v k 2 denotes the probability of occupation of a given pair state. Nevertheless, the pair operators do not satisfy exact bosonic commutation relations. In fact, for pair creation operators:
[ b k , b k ] 0 , where b k = c k c k
This deviates from canonical bosonic algebra, raising concerns about the true bosonic nature of these composite entities.
Another key issue arises from thermal perturbations. At finite temperatures, thermal excitations can break Cooper pairs. The critical temperature T c is set by the condition where the thermal energy k B T is comparable to the gap energy Δ . Taking T c 7.2 K for lead:
k B T c = ( 1.38 × 10 23 J / K ) × 7.2 9.94 × 10 23 J
This corresponds to 0.62 meV , which is in the same range as the superconducting gap. The pairing is thus extremely fragile, requiring only minute thermal agitation to disrupt the condensate. Unlike conventional bosons (such as atoms in Bose-Einstein condensates) that interact via strong external trapping potentials, Cooper pairs rely entirely on a delicate many-body ground state for their existence.
In addition, it is instructive to contrast the BCS condensation with Bose-Einstein condensation (BEC). In BEC, the critical temperature for an ideal gas of bosons is:
T c BEC = 2 π 2 k B m n ζ ( 3 / 2 ) 2 / 3
For a particle mass m 2 m e and density n 10 28 m 3 , this yields:
T c BEC 2 π ( 1.05 × 10 34 ) 2 1.38 × 10 23 × ( 2 × 9.1 × 10 31 ) 10 28 2.612 2 / 3 30 , 000 K
This estimate is orders of magnitude larger than observed superconducting transition temperatures. The disparity arises because Cooper pairs are not free bosons, and their condensation does not follow the BEC paradigm directly. Nonetheless, in the BCS-BEC crossover theory, where the pairing interaction is tuned from weak to strong coupling, one transitions from extended, overlapping Cooper pairs to tightly bound bosonic dimers. This suggests that the bosonic identity of Cooper pairs is strongly coupling-dependent.
Moreover, no boson-like conduction is observed in normal metals, even though electron-electron correlations and collective excitations exist. If electron pairs can exist in a condensed state at low temperatures, one would expect remnants of bosonic behavior at elevated temperatures or under specific conditions. However, phenomena such as superfluidity, second sound, or off-diagonal long-range order are absent in normal metallic states.
Experimental efforts to detect the bosonic character of Cooper pairs have relied on probing collective excitations. For instance, the phase mode (Anderson-Bogoliubov mode) is predicted as a Goldstone boson resulting from spontaneous U(1) symmetry breaking. However, in charged systems, this mode is lifted to the plasma frequency by the Anderson-Higgs mechanism [10]. This results in its absence from observable low-energy spectra, removing direct experimental access to the bosonic phase.
In conclusion, while Cooper pairs can be formally treated as bosons under specific conditions, their statistical behavior deviates from that of elementary bosons in both theory and observation. The lack of canonical commutation relations, the fragility of the pair state under thermal excitation, and the absence of bosonic phenomena in normal metals call into question the physical robustness of their bosonic identity. The fermionic nature of the constituents imposes intrinsic limitations on the coherence.

9. Lack of Universality in High- T c Superconductors

The discovery of high-temperature superconductors (HTS), beginning with cuprates in 1986, fundamentally challenged the universality of the BCS theory. While BCS theory explains conventional low-temperature superconductors with remarkable success, it fails to capture the phenomenology of HTS materials such as cuprates, iron pnictides, and heavy fermion systems. These classes of materials exhibit unconventional properties including non-Fermi liquid behavior, pseudogap phases, and anomalous scaling laws, which resist explanation within the weak-coupling, phonon-mediated framework of BCS.
One of the defining features of HTS materials is the presence of a pseudogap phase. This is a regime where a partial gap opens in the density of states above the superconducting transition temperature T c . The electronic excitation spectrum measured via angle-resolved photoemission spectroscopy (ARPES) [28] reveals that in cuprates such as Bi2Sr2CaCu2O8+δ, a gap-like feature persists up to T * > T c , which is not explained by BCS theory.
The presence of a pseudogap implies that pairing correlations exist without long-range phase coherence. In BCS, the energy gap Δ ( T ) closes continuously at T c :
Δ ( T ) = Δ 0 1 T T c
This functional form does not accommodate a gap that survives above T c , suggesting that the BCS order parameter is not applicable in the pseudogap regime.
Furthermore, isotope effect measurements yield anomalous results in HTS. In conventional superconductors, the isotope effect coefficient α defined by:
T c M α
typically satisfies α 0.5 , consistent with phonon-mediated pairing. However, in cuprates, α is often small or even negative [13], implying that phonons play a minimal role in the pairing mechanism. This contradicts the central tenet of BCS theory which relies on phonon interactions to induce an effective attractive potential between electrons.
Neutron scattering experiments provide further evidence against phonon-mediated mechanisms. In La2−xSrxCuO4, inelastic neutron scattering reveals a resonance peak at the antiferromagnetic wavevector, indicating the importance of spin fluctuations [14]. These fluctuations are strong in the underdoped region and may act as the pairing glue. The magnetic resonance energy E r scales with T c according to:
E r 5.8 k B T c
This scaling law, absent in conventional superconductors, supports the idea that magnetic excitations, not phonons, dominate in HTS.
Moreover, the d-wave symmetry of the order parameter in cuprates has been confirmed via phase-sensitive experiments [15], contrary to the s-wave symmetry predicted by BCS theory. The d-wave gap function has nodes along the Fermi surface, with the angular dependence:
Δ ( θ ) = Δ 0 cos ( 2 θ )
This gap symmetry leads to power-law behavior in thermodynamic observables such as the specific heat and penetration depth, differing markedly from the exponential suppression expected in s-wave BCS superconductors.
Iron-based superconductors further complicate the universality of BCS theory. These materials exhibit multiple Fermi surfaces and interband interactions, requiring theoretical frameworks beyond single-band BCS. In some iron pnictides, the pairing symmetry is thought to be s ± , where the gap changes sign between different Fermi pockets. Such a state can be stabilized by antiferromagnetic spin fluctuations rather than phonons [16].
Furthermore, scanning tunneling microscopy (STM) and spectroscopy reveal nanoscale electronic inhomogeneity in cuprates [17], with spatial variations in the gap magnitude on the order of nanometers. BCS theory assumes a homogeneous pairing potential, which is clearly violated in these systems. The presence of charge stripes and local electronic order demands a theory capable of describing real-space pairing mechanisms and dynamic correlation effects.
Heavy fermion superconductors such as CeCu2Si2 and UPt3 exhibit superconductivity near magnetic quantum critical points. These systems have extremely low carrier densities and large effective masses, with a characteristic Kondo lattice behavior. The pairing in these systems is likely mediated by critical spin fluctuations, entirely outside the BCS phonon framework [18].
In light of these facts, it is evident that BCS theory does not offer a universal explanation for superconductivity. While it remains valid for weak-coupling, low- T c materials with well-defined Fermi surfaces and phonon spectra, its assumptions are violated in the strongly correlated, magnetically active, and structurally complex environments of high- T c superconductors. The necessity for alternative pairing mechanisms—spin fluctuation, RVB states, or orbital fluctuations—reflects the inadequacy of BCS theory as a general theory of superconductivity.

10. No Direct Observation of Cooper Pairs

Despite being central to the BCS theory of superconductivity, Cooper pairs have never been directly observed as isolated entities. All evidence for their existence remains indirect, inferred from measurements of phase coherence, energy gaps in the electronic density of states, and macroscopic manifestations like the Meissner effect. This raises fundamental questions about whether superconductivity necessarily requires the concept of a bound electron pair.
In tunneling experiments such as those performed by Giaever [8], the differential conductance d I / d V reveals a gap in the density of states, with a coherence peak near the gap edge. The BCS theory explains this via the quasiparticle excitation spectrum:
E k = ( ϵ k μ ) 2 + Δ 2
The density of states N ( E ) in the superconducting state becomes:
N ( E ) = N 0 | E | E 2 Δ 2
for | E | > Δ , and zero within the gap. However, this observation only reflects the modified excitation spectrum and not the actual presence of bound electron pairs in real space.
Similarly, the Meissner effect — the expulsion of magnetic fields from a superconductor — is often interpreted as evidence of a macroscopic quantum state. The London equations,
j t = n s e 2 m E , × j = n s e 2 m B
predict an exponential decay of magnetic fields with penetration depth:
λ L = m μ 0 n s e 2
Again, this only implies the presence of a superconducting condensate with carrier density n s , and not that the carriers are explicitly paired electrons. Moreover, recent reinterpretations of superconductivity as a manifestation of topological order or quantum entanglement have challenged the necessity of a pair-based description.
From a measurement standpoint, no experimental setup has succeeded in capturing a Cooper pair as a discrete particle-like object. The composite nature of Cooper pairs suggests that they exist in momentum space and are delocalized over a coherence length ξ 100 nm. Imaging or trapping such an entity is beyond current spatial resolution and energy sensitivity of available probes.
Moreover, phase-sensitive Josephson junction experiments [19] are often cited as support for Cooper pairing due to the observed 2 e periodicity in the supercurrent. The current-phase relation in Josephson junctions is given by:
I = I c sin ( ϕ )
where ϕ is the phase difference between two superconducting wavefunctions. The flux quantum observed in superconducting rings is:
Φ 0 = h 2 e
Although this 2 e quantization is interpreted as reflecting Cooper pair charge, it is an inference based on flux quantization and interference patterns, rather than a direct observation of paired states.
Additionally, the observation of Andreev reflection [20], where an electron from a normal metal is reflected as a hole at the NS interface, is taken as indirect evidence of pairing. The retro-reflected hole accounts for a missing Cooper pair in the superconducting side, preserving charge conservation. The conductance at low bias is enhanced due to this process, and its temperature dependence tracks the superconducting gap. However, the analysis depends on boundary conditions and coherence.
Another line of argument relates to many-body wavefunction properties. The BCS state is characterized by long-range phase coherence and off-diagonal long-range order (ODLRO), a condition first formalized by Yang [21]. This condition is given by the asymptotic form of the two-particle density matrix:
ψ ( r ) ψ ( r ) ψ ( r ) ψ ( r ) | Φ ( r , r ) | 2 as | r r |
Yet, ODLRO describes a field correlation, not a measurement of bound states. It demonstrates that the state has coherent quantum correlations, which could arise from other mechanisms including spin-liquid ground states, entangled valence bonds, or topological field configurations.
Recent theories based on quantum entanglement [43] and emergent gauge fields suggest that superconductivity may emerge from entangled many-body ground states, without requiring discrete Cooper pair objects. These approaches emphasize global coherence, topological invariants, and entropic measures, instead of real-space bound pairs.
Furthermore, experiments on strongly disordered superconductors show fluctuation-dominated regimes where gap features persist even when global superconductivity vanishes [23]. This suggests that pairing and coherence may be separable phenomena, casting further doubt on whether pair formation is the fundamental essence of superconductivity or merely a convenient low-energy model.
In conclusion, all currently known evidence for Cooper pairs is indirect, relying on interpretation of collective electronic phenomena. Despite decades of research, there exists no imaging, trapping, or direct detection of a real-space, isolated Cooper pair. This opens the possibility that superconductivity could arise from other forms of quantum coherence or entanglement, challenging the necessity and universality of the Cooper pair concept.

11. Energetic Stability Puzzle

BCS theory posits that the superconducting state represents a lower energy configuration than the normal metallic state, and that this stabilization arises due to the formation of an energy gap Δ at the Fermi surface. However, a closer analysis reveals that the origin of this energy lowering is subtle and its physical interpretation remains conceptually opaque. The BCS ground state is a macroscopic quantum superposition involving millions of electrons.
The condensation energy, which represents the energy gained by entering the superconducting state, is derived from the BCS ground state energy. The energy difference per unit volume between the superconducting and normal states is given by:
U cond = E normal E BCS = 1 2 N ( 0 ) Δ 2
Here, N ( 0 ) is the density of states at the Fermi level per spin and Δ is the superconducting energy gap. For a typical superconductor such as lead with N ( 0 ) 10 47 J 1 m 3 and Δ 1.5 meV = 2.4 × 10 22 J , we compute:
U cond = 1 2 × 10 47 × ( 2.4 × 10 22 ) 2 2.9 × 10 3 J / m 3
This condensation energy is minuscule compared to the total electronic energy density in a metal, which is typically on the order of 10 8 J / m 3 . The energy stabilization provided by the superconducting state is thus only a tiny fraction of the system’s total energy budget, raising the question of how such a fragile energetic advantage can maintain coherence over macroscopic distances and timescales.
Furthermore, the critical magnetic field H c required to destroy superconductivity is related to the condensation energy via:
1 2 μ 0 H c 2 = U cond
Solving for H c using the condensation energy calculated above:
H c = 2 μ 0 U cond = 2 × 4 π × 10 7 × 2.9 × 10 3 0.068 T
This shows that a relatively small magnetic field of around 68 mT is sufficient to destroy the superconducting state in lead. This fragility further underscores the puzzle: if the state is truly a more stable energy configuration, why is it so easily disrupted by thermal fluctuations or external fields?
One partial explanation lies in the collective behavior of the electron system. The energy gap opens up due to an instability of the Fermi surface under attractive interactions. According to the BCS theory, the total energy is minimized by allowing electrons to pair in time-reversed states ( k , k ) . The pairing interaction in momentum space is modeled by an effective potential V acting within a small energy shell near the Fermi surface of width ω D :
Δ = 2 ω D exp 1 N ( 0 ) V
This expression for the energy gap shows that even a small attractive interaction V leads to an exponentially small gap. The exponential dependence highlights that the pairing is inherently weak and fragile. The gain in energy is offset by the extremely small value of Δ , which in turn governs the temperature and field scales that destroy superconductivity.
The mean-field nature of the BCS theory also contributes to this energetic puzzle. The BCS wavefunction is a coherent superposition that neglects fluctuations, which become increasingly important in low-dimensional systems or near the critical temperature T c . The Ginzburg-Levanyuk parameter:
G i = k B T c H c 2 ξ 3 2
quantifies the relative strength of thermal fluctuations. For small coherence length ξ and low H c , G i becomes large, indicating that the mean-field picture fails near T c . This means that the BCS energetic description does not fully account for fluctuation effects that can dominate the physics near the transition.
Another conceptual issue lies in the quantum coherence over macroscopic numbers of electrons. The BCS ground state involves a condensate of overlapping Cooper pairs, with a coherence length often exceeding hundreds of nanometers. It is unclear how such a highly correlated many-body wavefunction can spontaneously emerge from the disordered Fermi sea without an external organizing principle.
Furthermore, the fragile energetic advantage is especially puzzling when superconductivity persists in the presence of nonmagnetic disorder. According to Anderson’s theorem [7], the s-wave BCS state is robust against potential scattering. However, the condensation energy depends sensitively on the pairing amplitude, which should be degraded by spatial inhomogeneity. Yet in experiments, the critical temperature and superconducting properties remain remarkably stable.
In summary, the BCS explanation for the energetic stability of superconductivity presents a paradox. The energy gain per volume is extremely small, yet the state exhibits macroscopic quantum coherence. The fragility to magnetic and thermal perturbations suggests a shallow energy minimum, inconsistent with the robustness of phase coherence observed experimentally. These inconsistencies call for a deeper examination of the energetic landscape underlying superconductivity.

12. Inconsistencies with Superfluidity Analogies

Both superfluidity and superconductivity are manifestations of macroscopic quantum phenomena, exhibiting properties such as phase coherence, quantized vortices, and the emergence of a frictionless state. However, their microscopic interpretations differ markedly: while superconductivity is explained by the pairing of electrons into Cooper pairs within the BCS framework, superfluidity in systems like helium-4 is understood without invoking such two-body pairing mechanisms.
In the case of superfluid helium-4, the transition to a superfluid phase occurs below the lambda point T λ 2.17 K and is associated with the onset of Bose-Einstein condensation (BEC). The fraction of particles in the condensate at zero temperature is approximately n 0 0.1 n , indicating strong interaction effects [25]. The superfluid velocity v s is given by the gradient of a quantum phase:
v s = m ϕ
Here, m is the mass of a helium atom, and ϕ is the phase of the condensate wavefunction. The superfluid component has zero viscosity, which is directly observed in persistent flow and the absence of entropy transport.
In superconductors, the analogous expression for the supercurrent density is derived from the London equation, which assumes a charged condensate of density n s :
j s = n s e 2 m A
where A is the vector potential. Taking the curl of both sides and applying Maxwell’s equations yields the London penetration depth:
λ L = m μ 0 n s e 2
While both systems exhibit non-dissipative flow, the underlying mechanisms are different. In helium-4, no assumption is made about pairing; the system is bosonic in nature. This fundamental distinction leads to the question of whether superconductivity could be alternatively understood as a form of charged quantum fluid, without necessarily invoking the concept of Cooper pairing.
The hydrodynamic interpretation of superfluidity provides insight into such possibilities. The Landau two-fluid model separates the system into a normal component and a superfluid component, with total density:
ρ = ρ n + ρ s
The equations of motion follow from conservation laws and include the mutual friction term due to quantized vortices. The critical velocity in superfluid helium is derived from Landau’s criterion:
v c = min ϵ ( p ) p
where ϵ ( p ) is the excitation spectrum. For phonons, ϵ ( p ) = c p gives v c = c , the speed of sound. In superconductors, however, the analogous critical current density j c is determined by the depairing current:
j c n s e Δ k F
where k F is the Fermi wavevector. This derivation depends explicitly on the existence of a superconducting gap and Cooper pairs. However, if superconductivity were instead described by a charged quantum fluid model, a different route could be followed.
Phenomenological models based on quantum hydrodynamics have been proposed to describe superconductors in analogy with superfluids. The complex order parameter Ψ ( r , t ) obeys a nonlinear Schrödinger equation or Gross-Pitaevskii-type equation:
i Ψ t = 2 2 2 m + V eff + g | Ψ | 2 Ψ
Here, V eff is the electromagnetic potential, and g describes the interaction strength. This equation describes collective excitations and allows for the derivation of quantized vortices, Josephson effects, and critical velocities—all without invoking a microscopic pairing mechanism.
Notably, the quantization of circulation in superfluids:
v s · d l = h m n
has a counterpart in the quantization of magnetic flux in superconductors:
Φ = A · d l = h 2 e n
The difference by a factor of two is typically attributed to the 2 e charge of Cooper pairs. However, from a quantum fluid perspective, the flux quantization could arise from topological constraints on the phase of the macroscopic wavefunction, irrespective of pairing.
Furthermore, the observation of second sound—a temperature wave—exists in superfluid helium but not in conventional superconductors. However, second sound has been observed in high- T c superconductors under certain conditions [26], suggesting a possible extension of hydrodynamic analogies.
In conclusion, the phenomenological similarities between superfluidity and superconductivity—such as zero viscosity, quantized vortices, and phase coherence—raise the possibility that both phenomena could stem from a deeper underlying principle: macroscopic quantum order. The necessity of Cooper pairing in superconductivity might therefore be more a reflection of mathematical convenience within BCS theory than a physical imperative.

13. Alternative Theories Exist

The BCS theory of superconductivity, though remarkably successful for conventional low-temperature superconductors, faces critical limitations when applied to unconventional and high-temperature superconductors. As experimental discoveries in cuprates, iron pnictides, and topological materials accumulate, alternative theoretical frameworks have emerged that do not rely solely on phonon-mediated Cooper pairing. These frameworks include spin fluctuation-mediated pairing, and resonating valence bond (RVB).
In the spin fluctuation model, pairing is mediated not by lattice vibrations, but by antiferromagnetic spin fluctuations. This theory is particularly relevant to the cuprate high-temperature superconductors, where the parent compounds exhibit strong antiferromagnetic correlations. The effective interaction V eff in such a model takes the form:
V eff ( q , ω ) = 3 2 U ¯ 2 χ ( q , ω )
where U ¯ is an effective Hubbard interaction and χ ( q , ω ) is the dynamic spin susceptibility [27]. The pairing instability arises when χ is strongly peaked near a nesting vector such as ( π , π ) , leading to d-wave symmetry of the gap:
Δ ( k ) = Δ 0 cos k x cos k y
This anisotropic gap structure has been confirmed experimentally in angle-resolved photoemission spectroscopy (ARPES) measurements of the cuprates, thereby providing strong support for spin fluctuation models [28].
Another fundamentally different approach is the resonating valence bond (RVB) theory proposed by Anderson [3]. The RVB model views the superconducting state as a spin-liquid composed of dynamically resonating singlet bonds. This framework is naturally suited for materials near a Mott insulating state, and describes superconductivity as arising from the doping of a quantum spin liquid. The RVB wavefunction can be written as:
| RVB = { i , j } ϕ i j c i c j | 0
where ϕ i j encodes the bond amplitude. The emergent superconductivity is not from pairing due to an attractive force, but rather due to coherence of spin singlets.
In recent years, holographic superconductivity based on the AdS/CFT correspondence has also attracted attention. Within this approach, a strongly coupled quantum field theory in d dimensions is dual to a classical gravity theory in d + 1 dimensions. The basic model for holographic superconductors introduces a charged scalar field ψ in an AdS black hole background. The gravitational action includes:
S = d 4 x g R + 6 L 2 1 4 F μ ν F μ ν | ψ i A ψ | 2 m 2 | ψ | 2
When the black hole temperature falls below a critical value, the scalar field condenses near the horizon, leading to spontaneous symmetry breaking in the boundary theory—an analog of superconductivity [29]. While this theory does not yield quantitative predictions for materials, it demonstrates that superconducting-like order can emerge in non-BCS mechanisms rooted in strong coupling.
Topological and anyonic superconductivity models offer yet another departure from the Cooper pairing picture. These models involve exotic quasiparticles such as Majorana fermions that obey non-Abelian statistics. The Bogoliubov-de Gennes Hamiltonian in a topological superconductor takes the form:
H = ξ k Δ k Δ k * ξ k
Topological invariants, such as the Chern number or Z 2 index, characterize the ground state. The presence of zero-energy edge modes or vortex-bound Majorana states is a hallmark of this phase [30]. These features do not arise in conventional BCS theory and indicate a fundamentally different form of pairing or even non-pairing quantum order.
Collectively, these alternative theories highlight that the BCS concept of Cooper pairing may be neither sufficient nor necessary to describe all forms of superconductivity. In some cases, like in spin fluctuation or RVB theory, pairing emerges from non-phononic interactions. In others, such as holographic or topological models, the order parameter need not correspond to electron pairs at all. Instead, superconductivity could be viewed as an emergent phenomenon of quantum coherence, and topological order.

14. No Phonon Signature in Many Superconductors

The standard BCS theory attributes superconductivity to the formation of Cooper pairs mediated by electron-phonon interactions. This mechanism predicts that phonons should play a central role in the superconducting transition. Consequently, a critical test of the phonon-mediated pairing hypothesis lies in direct observation of phonon anomalies—such as softening, linewidth broadening, or mode splitting—near the superconducting transition temperature T c .
According to the Migdal-Eliashberg theory, which extends BCS to include strong-coupling effects, the self-energy correction of electrons due to phonon interactions is expressed through the Eliashberg function α 2 F ( ω ) . The phonon-mediated electron-electron coupling constant λ is given by:
λ = 2 0 α 2 F ( ω ) ω d ω
The transition temperature T c is approximately given by the McMillan formula:
T c = Θ D 1.45 exp 1.04 ( 1 + λ ) λ μ * ( 1 + 0.62 λ )
where Θ D is the Debye temperature and μ * is the Coulomb pseudopotential. If λ is large, strong phonon signatures should be detectable through spectroscopic techniques.
Yet, neutron scattering and Raman spectroscopy in cuprate superconductors, which exhibit T c values exceeding 90 K, reveal no significant changes in phonon dispersion or linewidths near T c [31]. For example, in optimally doped YBa2Cu3O7−δ, extensive neutron inelastic scattering data show that most phonon modes remain unchanged across the superconducting transition, despite the presence of a well-defined gap in the electronic density of states.
Moreover, isotope effect measurements, a cornerstone of BCS theory, are markedly suppressed or even absent in high- T c superconductors. The canonical isotope effect index α is defined by:
α = d ln T c d ln M
where M is the ionic mass. In classic superconductors like Pb or Hg, α 0.5 , consistent with phonon involvement. However, in La2−xSrxCuO4 or Bi2Sr2CaCu2O8, measured values of α are anomalously small or even negative [2]. Such deviations indicate that phonons may not be central to the pairing mechanism in these systems.
The absence of phonon anomalies can also be analyzed through phonon self-energy corrections. The real part of the phonon self-energy Σ ( ω ) leads to mode renormalization (softening or hardening), while the imaginary part Im Σ ( ω ) contributes to broadening:
Δ ω = Re Σ ( ω ) , Γ = 2 Im Σ ( ω )
In materials like MgB2, where the BCS mechanism is accepted, significant phonon softening near T c is observed in the E 2 g mode due to strong electron-phonon coupling [32]. However, similar features are conspicuously absent in high- T c cuprates and iron pnictides, where the superconducting transition leaves most phonon branches unaffected.
Additionally, first-principles calculations of phonon dispersions using density functional perturbation theory (DFPT) often show that electron-phonon coupling in cuprates and Fe-based superconductors is insufficient to account for the high observed T c . The total calculated λ values in these families remain well below 0.5, placing them outside the strong coupling regime required for high- T c superconductivity via phonons [33].
These discrepancies suggest that mechanisms other than electron-phonon interaction are likely responsible for the pairing in many unconventional superconductors. Alternative bosonic modes, such as magnetic excitations or spin fluctuations, may be responsible for the observed superconductivity. This scenario is further supported by the strong temperature and doping dependence of magnetic susceptibility and the resonance peaks observed in inelastic neutron scattering in many cuprates [34].
In summary, the lack of phonon anomalies in a broad class of superconductors challenges the universality of phonon-mediated pairing as proposed by BCS theory. The absence of mode softening, isotope effect suppression, and inconsistency with first-principles predictions collectively indicate that other mechanisms may be at play in enabling high-temperature superconductivity. While phonons may still influence some properties, they do not appear to drive the pairing transition in these complex systems.

15. Topological Superconductivity and Its Implications for the Cooper Pair Paradigm

The theory of topological superconductivity provides a compelling framework that diverges fundamentally from the traditional phonon-mediated Bardeen-Cooper-Schrieffer (BCS) theory. In contrast to the weak coupling, electron-phonon attraction picture that defines BCS superconductivity, topological superconductors are characterized by a nontrivial topology in their quasiparticle wavefunctions and exhibit exotic edge states such as Majorana zero modes.
In conventional BCS theory, the pairing Hamiltonian is based on spin-singlet, s-wave symmetry, leading to an energy gap that is isotropic in momentum space. However, in topological superconductors, the pairing symmetry is typically p-wave or higher angular momentum, and the gap structure may be nodal or anisotropic. For instance, the Bogoliubov-de Gennes (BdG) Hamiltonian for a spinless p-wave superconductor in one dimension, such as the Kitaev chain, is written as:
H BdG = k ξ k c k c k + Δ k c k c k + Δ k * c k c k
where ξ k = 2 t cos ( k ) μ is the kinetic energy term, t is the hopping amplitude, μ is the chemical potential, and Δ k = Δ sin ( k ) represents the p-wave pairing potential [30].
The topological invariant for such systems can be computed using the winding number or the Pfaffian of the Hamiltonian matrix. For example, in the Kitaev chain, the winding number ν is defined as:
ν = 1 2 π π π d k d d k arg [ Δ k + i ξ k ]
This invariant takes values in { 0 , 1 } , distinguishing topologically trivial and nontrivial phases. The phase transition occurs when the gap closes, i.e., Δ k = 0 and ξ k = 0 , which happens at μ = ± 2 t .
Unlike the Cooper pair condensate, topological superconductors can host unpaired zero-energy edge states known as Majorana bound states. These states satisfy the self-conjugacy condition:
γ = γ
This property allows them to obey non-Abelian statistics, making them fundamentally distinct from BCS quasiparticles and strongly supporting the idea that superconductivity need not rely on Cooper pairing as a physical mechanism [35].
Another striking contrast lies in the robustness of topological superconductivity. The edge states are protected by the topology of the bulk band structure and cannot be removed by any local perturbation that does not close the energy gap. In this sense, superconducting coherence arises from topological constraints rather than delicate electron-phonon interactions. This kind of robustness is absent in conventional BCS theory, where small perturbations can easily disrupt the coherence of Cooper pairs.
Furthermore, the formation of a superconducting gap in topological superconductors does not always correlate with phonon modes or isotope effects. For instance, materials such as FeTe1−xSex, which are candidates for topological superconductivity, exhibit extremely weak electron-phonon coupling. In these systems, the estimated coupling constant λ is less than 0.2, well below the threshold required to support phonon-mediated pairing in the McMillan formula:
T c = Θ D 1.45 exp 1.04 ( 1 + λ ) λ μ * ( 1 + 0.62 λ )
If λ 1 , then T c should be negligible. Yet, these materials exhibit T c values above 10 K, further suggesting a non-phononic origin of the superconducting state [33].
Topological superconductors also display quantized conductance plateaus and thermal transport signatures consistent with chiral edge modes. In particular, the quantized thermal Hall conductance κ x y / T is predicted to follow:
κ x y T = π 2 k B 2 3 h C
where C is the Chern number, k B is Boltzmann’s constant, and h is Planck’s constant. This quantization has no analog in BCS theory and originates solely from the topological properties of the bulk gap function [36].
In summary, the theory of topological superconductivity provides a fundamentally distinct and mathematically rigorous alternative to the BCS framework. The phenomena of Majorana edge states, topological invariants, nontrivial pairing symmetries, and the absence of phonon signatures collectively indicate that Cooper pairing is neither necessary nor sufficient for superconductivity in all materials. Instead, superconductivity may emerge from the topology of the many-body wavefunction itself.

16. Fundamental Issues with Cooper Pair Formalism in Low-Temperature Superconductors

While the Bardeen-Cooper-Schrieffer (BCS) theory of superconductivity has been widely successful in describing many features of low-temperature superconductors (LTS), there remain profound foundational issues that challenge the literal interpretation of the Cooper pair formalism. These issues persist even in canonical superconductors such as aluminum, lead, tin, and mercury. In this section, we outline five key areas of concern, grounded in precise calculations and observable discrepancies.

16.1. Drift Velocity vs. Electromagnetic Signal Propagation

The observed current in superconductors is carried by electrons with an extremely low drift velocity. The drift velocity v d is given by:
v d = I n e A
For a current I = 1 A , charge carrier density n 8.5 × 10 28 m 3 , cross-sectional area A = 1 mm 2 , and elementary charge e = 1.6 × 10 19 C , we find:
v d 1 8.5 × 10 28 · 1.6 × 10 19 · 10 6 7.4 × 10 4 m / s
This drift velocity is in stark contrast to the speed at which electrical signals propagate in the superconducting medium, which is governed by the group velocity of electromagnetic waves and is on the order of 10 8 m / s . The existence of superconductivity implies macroscopic quantum coherence rather than transport by particle motion. The low drift velocity undermines the narrative that Cooper pairs physically “flow” as charge carriers [39].

16.2. Delocalization and the Coherence Length

The Cooper pair is often portrayed as a bound state of two electrons, but in BCS theory, this “pair” is a delocalized object with an extremely large spatial extent. The BCS coherence length ξ 0 is given by:
ξ 0 = v F π Δ
For low-temperature superconductors, typical parameters are v F 10 6 m / s , and superconducting gap Δ 1 meV = 1.6 × 10 22 J , yielding:
ξ 0 1.05 × 10 34 · 10 6 π · 1.6 × 10 22 21 nm
This coherence length is approximately 100 times the interatomic spacing, indicating that the pair is not a localized two-electron entity but a nonlocal wave-like excitation. Such extreme delocalization raises questions about the physical validity of the “bound electron pair” interpretation, even in conventional superconductors [40].

16.3. Energetic Fragility of the Superconducting State

If Cooper pairs are energetically stable bound states, their destruction should require energy comparable to their binding energy. However, superconductivity in LTS is destroyed by minute perturbations. For example, lead loses superconductivity above a critical magnetic field H c 0.08 T , corresponding to a thermodynamic critical energy density:
U c = H c 2 2 μ 0
Substituting μ 0 = 4 π × 10 7 N / A 2 , we obtain:
U c ( 0.08 ) 2 2 · 4 π × 10 7 2.5 × 10 3 J / m 3
This energy density is negligible compared to the Fermi energy density, which is of the order 10 9 J / m 3 . The fact that a microscopic fraction of energy can destroy superconductivity suggests that the superconducting state is not a stable bound state of electron pairs but a fragile macroscopic quantum state vulnerable to decoherence [37].

16.4. Absence of Direct Cooper Pair Detection

Even in low-temperature superconductors, all evidence for Cooper pairs is indirect. Techniques such as tunneling spectroscopy, specific heat measurements, and electromagnetic response show the presence of a gap or phase coherence but do not directly detect real-space electron pairs. The lack of direct detection calls into question the ontological status of Cooper pairs. If such entities exist as real-space objects, they should be observable under scanning tunneling microscopy.

16.5. Misinterpretation of the BCS Ground State

The BCS ground state is a coherent superposition of particle pairs in momentum space. The wavefunction is given by:
| Ψ BCS = k u k + v k c k c k | 0
This is not a real-space bound state of two electrons but a macroscopic entangled state with no direct analog in real-space physics. The operators c k σ create electrons with well-defined momentum and spin. Therefore, interpreting the BCS state as a condensate of spatially bound electron pairs is conceptually misleading. It is more accurate to view the state as a Fock-space condensate with off-diagonal long-range order (ODLRO), a concept introduced by Yang [21].

17. Philosophical and Interpretational Critique of the BCS Formalism

The Bardeen-Cooper-Schrieffer (BCS) theory has become the paradigmatic framework for understanding conventional superconductivity. While its mathematical structure yields predictions that align with several key experimental results, including the energy gap, critical temperature, and isotope effect, the theory carries philosophical and interpretational ambiguities that demand critical scrutiny. In this section, we explore the disconnect between the formalism and physical ontology.

17.1. Formal vs. Physical Ontology of the BCS State

The BCS ground state is a variational ansatz composed of entangled fermionic states in momentum space:
| Ψ BCS = k u k + v k c k c k | 0
This wavefunction exhibits off-diagonal long-range order (ODLRO) in momentum space, but it has no clear real-space interpretation as a set of spatially localized bound pairs. The coefficients u k and v k satisfy:
| u k | 2 + | v k | 2 = 1
The probability amplitudes v k 2 correspond to occupation probabilities of momentum states, not real-space electron configurations. Thus, while the formalism predicts macroscopic coherence and energy gaps, it does not justify the conceptual leap to the picture of real-space “Cooper pairs” that traverse the lattice as composite bosons.

17.2. Interpretational Conflicts: Bound State or Emergent Coherence?

The Cooper problem shows that two electrons above a filled Fermi sea can form a bound state via a weak attractive interaction, with binding energy:
E bind = 2 ω D e 2 / N ( 0 ) V
where ω D is the Debye frequency, N ( 0 ) the density of states, and V the effective electron-phonon interaction. However, this solution is derived for a pair of non-interacting electrons, not a thermodynamically large system. When extended to a many-body system, the BCS energy gap is derived variationally, rather than as an eigenstate of a known physical interaction Hamiltonian:
Δ = ω D exp 1 N ( 0 ) V
The absence of an explicit real-space interaction potential producing this gap, particularly over the coherence length ξ 0 10 100 nm , renders the physical interpretation ambiguous. The so-called "binding" does not emerge from a two-body interaction but from a mean-field-like condensation in Fock space.

17.3. Problems with Particle Number Non-Conservation

The BCS wavefunction is not an eigenstate of particle number, and its formulation breaks gauge symmetry explicitly. This has been rationalized as a feature of spontaneous symmetry breaking in a macroscopic quantum state. However, the use of non-conserving particle number states creates interpretational dilemmas regarding what constitutes a “pair”. Particle number restoring projections are rarely invoked in practical calculations. The standard BCS state,
| Ψ BCS = N C N | N particles
is a superposition over different total electron numbers. Thus, the “pairing” is not a real configuration of two-electron states but a fluctuation in many-body occupancy that gives rise to anomalous expectation values such as:
c k c k = u k v k 0
This anomalous average signals long-range order but does not correspond to an actual measurement outcome of a two-electron bound state. The expectation value is formally defined, but its operational meaning is unclear without a detector sensitive to such pair amplitudes [38].

17.4. Microscopic Hamiltonian and Emergence Issues

The BCS Hamiltonian is based on a reduced pairing interaction of the form:
H = k σ ϵ k c k σ c k σ k , k V k , k c k c k c k c k
Here, V k , k is a phenomenological input, justified post hoc via the electron-phonon mechanism. However, no first-principles derivation of this interaction from the underlying Coulomb + lattice Hamiltonian exists. The phonon mechanism gives corrections to the dielectric function, but it cannot easily account for attraction over length scales ξ 0 a , where a is the lattice constant. The emergent pair interaction appears to be a projection, not a fundamental force.

17.5. Conceptual Fragility and Misleading Analogies

The common analogy of Cooper pairs as “electron molecules” or composite bosons moving without resistance obscures the actual structure of the theory. The current in BCS is calculated from the phase gradient of the order parameter:
J s = 2 e n s m ϕ
where n s is the superfluid density and ϕ is the global phase. This equation suggests that the transport is a phase effect, not motion of discrete composite particles. Furthermore, if Cooper pairs were real-space objects, they would be subject to local decoherence, collisions, and scattering, yet none are observed. The BCS theory avoids this by never describing localized pair motion — thus, the composite boson analogy is metaphorical, not physical [39].

17.6. Summary of the Ontological Inconsistencies

The BCS theory provides a successful formalism but fails to deliver a physically consistent, ontologically clear explanation of superconductivity. Cooper pairs do not exist as real-space bound particles; the energy gap is not produced by a tangible interaction; the wavefunction does not conserve particle number; and current is carried by phase gradients, not quasiparticle flow. These contradictions imply that BCS is best understood as an effective theory of emergent order — a phenomenological success.

18. Experimental Observations in Low-Temperature Superconductors: Data and Interpretational Challenges

While the BCS theory quantitatively matches a variety of experimental signatures in low-temperature superconductors (LTS), the interpretation of these measurements remains open to question. In this section, we present representative experimental data from tunneling spectroscopy, specific heat, and magnetic critical fields in canonical LTS such as lead (Pb), tin (Sn), and aluminum (Al). We examine whether the data necessitate the Cooper pair framework or if alternative interpretations are viable.

18.1. Tunneling Spectroscopy and Energy Gap Measurements

Tunneling spectroscopy allows for direct observation of the superconducting energy gap Δ through current-voltage characteristics. The differential conductance d I d V is proportional to the density of states N ( E ) , given by:
N ( E ) = N 0 | E | E 2 Δ 2 , | E | > Δ
This singularity at E = Δ produces a characteristic peak in the conductance. For lead, measurements yield Δ Pb 1.35 meV at T = 0 K , and a critical temperature T c 7.2 K , giving the dimensionless ratio:
2 Δ k B T c 2 · 1.35 × 10 3 1.38 × 10 23 · 7.2 4.3
This exceeds the canonical BCS value 3.52 , suggesting that strong-coupling effects or deviations from the weak-coupling BCS regime may be at play [8]. For tin (Sn), Δ 0.6 meV , T c = 3.7 K , again yielding 2 Δ / k B T c > 3.52 .

18.2. Specific Heat Discontinuity at T c

In BCS theory, the specific heat in the superconducting state exhibits a discontinuous jump at T = T c , expressed as:
C s C n C n | T c = 1.43
where C s and C n are the specific heats in the superconducting and normal states, respectively. Experimental data for aluminum (Al) and tin (Sn) agree with this prediction within experimental uncertainty [37]. However, the behavior at T T c deviates from exponential suppression in some LTS, where residual states persist:
C s ( T ) exp Δ k B T , T T c
The presence of sub-gap states or impurity-induced broadening undermines the simple BCS picture, implying more complex many-body interactions than a clean, mean-field transition.

18.3. Magnetic Critical Field and Thermodynamic Energy

The thermodynamic critical field H c defines the field strength at which superconductivity is destroyed. The associated condensation energy density U c is given by:
U c = H c 2 2 μ 0
For lead with H c ( 0 ) = 0.08 T , and μ 0 = 4 π × 10 7 H / m , the condensation energy is:
U c ( 0.08 ) 2 2 · 4 π × 10 7 2.5 × 10 3 J / m 3
This energy density is minuscule compared to the normal-state Fermi energy density 10 9 J / m 3 . Therefore, the superconducting state represents an extremely delicate energy minimization, raising concerns about how such a small gain can stabilize a macroscopic quantum state without invoking collective coherence beyond pairing.

18.4. London Penetration Depth and Superfluid Density

The London penetration depth λ L characterizes the exponential decay of magnetic fields into the superconductor:
λ L = m μ 0 n s e 2
Measurements for aluminum yield λ L 50 nm , with a carrier density n s 1.8 × 10 29 m 3 . From this, one can infer the effective mass or superfluid density. The fact that these parameters are sensitive to disorder and vary between samples casts doubt on the universality of BCS expressions in real materials [39].

18.5. Interpretational Gaps in Experimental Agreement

Although the BCS theory can be tuned to fit many experimental curves in LTS, the existence of sub-gap states, deviations in the 2 Δ / k B T c ratio, and residual specific heat suggest that the match is not universal or complete. Furthermore, there is no direct evidence for the real-space existence of Cooper pairs or their dynamics. All observables — energy gap, Meissner effect, and heat capacity — can potentially be modeled via alternative frameworks invoking macroscopic coherence.

19. Mismatch Between Drift Velocity and Signal Propagation Speed in Superconductors

One of the often-overlooked inconsistencies in the microscopic interpretation of superconductivity, even within the conventional low-temperature regime, is the striking mismatch between the actual electron drift velocity and the observed speed at which electrical signals propagate through a superconducting material. In classical electrodynamics, the drift velocity of electrons in a conductor is defined by the equation:
v d = I n e A
where I is the current, n is the electron density, e is the elementary charge, and A is the cross-sectional area of the conductor. Let us evaluate this in the case of lead (Pb), a canonical low-temperature superconductor, at liquid helium temperature ( T = 4.2 K ).

19.1. Quantitative Estimate of Drift Velocity

Taking I = 1 A , n 10 29 m 3 , A = 1 mm 2 = 1 × 10 6 m 2 , and e = 1.6 × 10 19 C , the drift velocity becomes:
v d = 1 ( 10 29 ) ( 1.6 × 10 19 ) ( 1 × 10 6 ) 6.25 × 10 5 m / s
This velocity is exceedingly small. For comparison, the speed of sound in solids is on the order of 10 3 m / s , and electromagnetic signals travel at a significant fraction of the speed of light ( c 3 × 10 8 m / s ) in a metal due to the plasma frequency response. Despite this, current in a superconductor establishes almost instantaneously over macroscopic distances, indicating that the transmission mechanism is not due to physical transport of electrons but rather a collective quantum effect.

19.2. Collective Phase Coherence and Signal Velocity

In BCS theory, the superconducting current arises from a macroscopic wavefunction Ψ ( r ) = n s e i ϕ ( r ) , where n s is the superfluid density. The current density is then expressed via the London equation as:
J s = 2 e n s m ϕ
This current does not stem from individual electron drift but from the gradient of a quantum phase field. The establishment of current, therefore, is limited by the speed of phase propagation, not particle velocity. Empirically, switching times in superconducting circuits such as SQUIDs are on the order of nanoseconds or less, suggesting that signal propagation occurs at nearly relativistic speeds.

19.3. Implications for the Cooper Pair Paradigm

Cooper pairs are extended objects with coherence lengths:
ξ 0 = v F π Δ
Using v F 10 6 m / s and Δ 1 meV , we obtain:
ξ 0 1.05 × 10 34 · 10 6 π · 1.6 × 10 22 20 nm
This implies that Cooper pairs extend over many lattice constants. If one assumes a real-space pair picture, the slow drift velocity of the electron constituents should lead to internal friction, scattering, or viscous effects. However, no such dissipation is observed. Even when a supercurrent flows, the kinetic energy per unit volume is:
E k = 1 2 n s m v d 2 1 2 · 10 29 · 9.1 × 10 31 · ( 10 4 ) 2 4.5 × 10 10 J / m 3
This energy is negligible, suggesting that macroscopic current is not carried by particle motion but rather by a non-dissipative phase field. This challenges the notion that Cooper pairs, as real-space entities, mediate current through center-of-mass motion.

19.4. Alternative Interpretation: Quantum Hydrodynamics

A more appropriate analogy may be drawn from quantum hydrodynamics, where the phase field of the condensate governs the macroscopic flow. In superfluid helium, flow is similarly irrotational and non-dissipative, and the speed of disturbance is set by the speed of sound, not atom drift. If superconductivity is fundamentally a phase-coherent ground state, then signal propagation can be instantaneous relative to particle motion. Therefore, the emphasis on Cooper pair kinetics may obscure the true origin of transport.

19.5. Conclusion

The enormous disparity between drift velocity and signal speed in low-temperature superconductors cannot be reconciled with a literal interpretation of Cooper pair motion. Instead, it implies that superconductivity is a manifestation of a global phase coherence, and that current is a field-level phenomenon arising from the quantum mechanical phase of the system. This holds even in conventional superconductors, and thus poses a fundamental interpretational challenge to the conventional BCS-Cooper pair ontology.

20. Lack of Real-Space Localization of Cooper Pairs in Low-Temperature Superconductors

The foundational mechanism of superconductivity in the BCS framework is often described via the formation of Cooper pairs—pairs of electrons with opposite momentum and spin, which condense into a macroscopic quantum state. While this picture operates rigorously in momentum space, its interpretation in real space introduces conceptual inconsistencies, particularly in low-temperature superconductors (LTS). The real-space extent of Cooper pairs, measured by the coherence length ξ 0 , highlights a severe shortcoming.

20.1. Coherence Length and Spatial Delocalization

The BCS coherence length in clean systems is given by the expression:
ξ 0 = v F π Δ
where v F is the Fermi velocity and Δ is the superconducting energy gap. In typical LTS such as lead (Pb) or tin (Sn), we may take v F 10 6 m / s and Δ 1 meV = 1.6 × 10 22 J . Substituting into the above equation gives:
ξ 0 = 1.05 × 10 34 · 10 6 π · 1.6 × 10 22 20 × 10 9 m = 20 nm
This coherence length is approximately two orders of magnitude larger than the interatomic spacing in metals, which is typically a 0.2 nm . Thus, each Cooper pair encompasses roughly N ( ξ 0 / a ) 3 ( 100 ) 3 = 10 6 atoms, indicating that the pair is not localized but delocalized over an enormous region in real space.

20.2. Interpretational Issues with Bound Pair Picture

Given that Cooper pairs are not confined to spatially adjacent electrons, the traditional picture of pairing—evoking a hydrogen-like bond or proximity-induced coupling—is misleading. The pairing occurs in momentum space, where electrons of opposite momentum ( k and k ) form time-reversed states. The resulting wavefunction, derived from BCS theory, is a superposition of such pairs:
| Ψ BCS = k u k + v k c k c k | 0
The real-space pair correlation function is obtained via Fourier transform:
Ψ ( r ) = k v k u k e i k · r
For small gaps, this wavefunction decays slowly in space, explaining the large coherence length. However, this spatial extension conflicts with any interpretation that views Cooper pairs as quasi-localized bosonic entities.

20.3. Consequences for Observables and Theory

The delocalized nature of Cooper pairs raises several questions. First, if these are true bound states, why is there no spectroscopic signature of individual pair excitations in the tunneling or Raman data? Second, how can such extended entities exhibit no scattering or decoherence, even in disordered systems? Third, what role does the real-space structure play in observable superconducting parameters such as critical current or magnetic field penetration depth?
These inconsistencies become particularly stark when considering type-I LTS such as aluminum and lead, where the superconducting state is highly homogeneous and bulk-like. In such cases, one expects that disorder or inhomogeneity should lead to phase decoherence over lengths much smaller than ξ 0 , yet phase coherence is preserved over macroscopic distances.

20.4. Delocalization as a Challenge to Real-Space Models

Attempts to construct real-space models of Cooper pairs (e.g., bipolaron or lattice-coupled models) fail to capture the observed coherence length or the gap-to- T c ratio. Furthermore, the assumption of bosonic condensation of preformed pairs is inconsistent with the measured critical exponents and specific heat behavior near T c [37]. The lack of any observable short-range real-space pair suggests that pairing should not be interpreted as a localized bond.

20.5. Conclusion

The fact that Cooper pairs in low-temperature superconductors extend over tens of nanometers and include hundreds of thousands of atoms contradicts the simplistic bound pair picture often presented. The physical nature of pairing is momentum-space delocalization, not real-space coupling, and models relying on a real-space picture obscure rather than clarify the quantum mechanical origin of superconductivity. Even in low-temperature superconductors, the lack of real-space localization challenges the interpretation.

21. Fragility of the Superconducting State in Low-Temperature Superconductors

Despite the theoretical elegance and predictive successes of the BCS theory, a closer inspection of experimental data reveals a paradox: the superconducting state, although energetically favorable, is remarkably fragile. In low-temperature superconductors (LTS), this state can be destroyed by modest magnetic fields, low-current injections, or minimal thermal excitations. This fragility stands in contrast to the large energy scales of the normal state, suggesting a reevaluation of what "stability" means.

21.1. Energy Scales: A Quantitative Comparison

Let us first examine the relevant energy scales. The Fermi energy E F , which characterizes the kinetic energy of conduction electrons, is typically on the order of:
E F 1 10 eV
In contrast, the superconducting energy gap Δ , even in the cleanest LTS such as Pb or Al, is:
Δ 1 meV = 1.6 × 10 22 J
Thus,
E F Δ 10 3 10 4
This disparity raises a conceptual issue. If superconductivity is the result of a minor rearrangement near the Fermi surface, why is the ground state so sensitive to perturbations that are three or four orders of magnitude smaller than the electronic bandwidth?

21.2. Sensitivity to Magnetic Fields

Superconductivity in type-I LTS is suppressed by magnetic fields as low as:
H c 0.01 T
The magnetic energy density corresponding to this field is:
E B = B 2 2 μ 0 = ( 0.01 ) 2 2 × 4 π × 10 7 4 × 10 2 J / m 3
In contrast, the condensation energy per unit volume for typical LTS is on the order of:
E cond = 1 2 N ( 0 ) Δ 2
Assuming N ( 0 ) 10 47 J 1 m 3 and Δ 10 3 eV , we get:
E cond 1 2 · 10 47 · ( 1.6 × 10 22 ) 2 2 × 10 5 J / m 3
Hence, a small magnetic field can inject enough energy to exceed the condensation energy and destroy the superconducting state.

21.3. Thermal Fragility

The energy of thermal excitations is given by k B T . Even at T = 4.2 K , the thermal energy is:
k B T 1.38 × 10 23 · 4.2 5.8 × 10 23 J 0.36 meV
This is already a significant fraction of Δ . Therefore, even minute thermal fluctuations can break Cooper pairs, suggesting a finely balanced state that lacks robustness.

21.4. Critical Currents and Depairing Effects

Injected currents also destroy superconductivity above a critical threshold. The depairing current density J c is given by:
J c = 2 e n s Δ k F
For typical values n s 10 28 m 3 , k F 10 10 m 1 , and Δ 1 meV , one finds that J c 10 6 A / m 2 , which is small on macroscopic scales.

21.5. Interpretational Challenge

The BCS model claims that the superconducting state is a ground state of the electron system, stabilized by an energy gap and protected by long-range coherence. However, this state appears to be more accurately described as a delicate balance of quantum coherence rather than a truly robust minimum of energy. The ease with which it is disrupted implies that it is highly sensitive to boundary conditions, topology, and macroscopic phase alignment, rather than microscopic pair binding alone [39].

21.6. Conclusion

The fragility of superconductivity in LTS contradicts the notion of a globally stable, energetically robust ground state predicted by BCS. The state is disrupted by perturbations many orders of magnitude smaller than the Fermi energy. This indicates that the superconducting state may rely more on macroscopic phase coherence and less on local binding energy. These observations call for a deeper interpretation of what stabilizes the superconducting state and challenge the completeness of the Cooper-pair-based theory.

22. No Direct Detection of Cooper Pairs — Even in Low-Temperature Superconductors

The Cooper pair is the foundational construct of the BCS theory, envisioned as a weakly bound state of two electrons with opposite momentum and spin. Despite decades of intensive experimental research, particularly in low-temperature superconductors (LTS) such as aluminum (Al), tin (Sn), and lead (Pb), no direct measurement of a Cooper pair as a real-space, two-particle entity has been made. All available evidence remains phenomenological, macroscopic, and indirect.

22.1. Tunneling Spectroscopy and Energy Gap

Tunneling spectroscopy in superconducting junctions is perhaps the most cited empirical success of BCS theory. The presence of an energy gap Δ manifests as a suppression of conductance at low voltages and coherence peaks at e V = ± Δ . The superconducting density of states N s ( E ) follows:
N s ( E ) = N 0 | E | E 2 Δ 2 for | E | > Δ
This equation correctly describes experimental results in LTS materials [39]. However, this gap pertains to quasiparticle excitation and not to a direct two-particle bound state. The tunneling current in a superconductor-insulator-normal metal (SIN) junction involves single-electron processes and does not spatially resolve any internal structure of the condensate.

22.2. Meissner Effect and Macroscopic Coherence

The Meissner effect, described by the London penetration depth λ L , is explained via the London equations. The superconducting state expels magnetic fields due to macroscopic coherence of the wavefunction, as seen from:
J s = n s e 2 m A
This implies:
2 B = 1 λ L 2 B , λ L = m μ 0 n s e 2
This derivation contains no reference to two-particle states. The field expulsion depends on the phase stiffness of the macroscopic wavefunction rather than on direct evidence of pair formation.

22.3. Josephson Effect and Absence of Pair Signature

In a Josephson junction, the current–phase relationship follows:
I = I c sin ( ϕ )
where ϕ is the phase difference across the junction. The associated energy:
E J = I c 2 e
reflects the energy stored in the junction due to coherent phase difference. While this phenomenon is described in terms of Cooper pair tunneling, what is actually observed is a macroscopic phase-dependent current, not a measurement of a two-electron bound state. No spectral line or spatially localized image of a Cooper pair has emerged from Josephson measurements [38].

22.4. Scanning Probe and Spectroscopy Absence

Modern techniques such as scanning tunneling microscopy (STM) have sub-nanometer resolution, and yet no Cooper pair has ever been imaged as a spatially correlated electron pair. STM measures the local density of states (LDOS), and coherence peaks in the LDOS appear at the gap edges, but do not reveal the internal structure of the condensate:
LDOS ( E ) Im G ( E , r , r )
where G is the Green’s function of the system. Even high-resolution STM on Pb and Nb has not reported any localized bound-state features that can be ascribed to real-space Cooper pairing [37].

22.5. Absence in Inelastic Spectroscopy

Inelastic neutron scattering and Raman spectroscopy are capable of identifying sharp excitations and collective modes. In LTS, these tools observe phonon softening and collective excitations, but no peak or mode identifiable as a direct excitation of a Cooper pair. If a bound state of two electrons existed with internal structure, it should have a characteristic excitation frequency or spectroscopic fingerprint. The absence of such a feature reinforces the notion that Cooper pairs are a theoretical device.

22.6. Conclusion

The lack of direct detection of Cooper pairs, even in ideal low-temperature systems, highlights a deep gap between the formalism and the empirical world. The Cooper pair remains an abstraction constructed to explain collective behavior, but not a directly observed object. Despite extensive experimental access and high-resolution probes, the failure to image, isolate, or spectroscopically identify a real-space Cooper pair calls into question its ontological status in the physical theory of superconductivity.

23. BCS Ground State Is Not a Paired State in Position Space

The BCS theory describes the superconducting ground state through a coherent superposition of electron pairs with opposite momentum and spin. The canonical form of the BCS wavefunction is:
| Ψ BCS = k ( u k + v k c k c k ) | 0
This state exists in momentum (Fock) space and represents a condensate of overlapping momentum-entangled electron pairs. Importantly, it is not a bound state in position space like a diatomic molecule. The pair amplitude:
Ψ BCS | c r 1 c r 2 | 0
is nonzero for all separations | r 1 r 2 | , decaying slowly over the coherence length ξ 0 , which is typically tens to hundreds of nanometers in low-temperature superconductors (LTS). This implies delocalized pairing, not a spatially bound state [39].

23.1. Coherence Length and Delocalization

The coherence length ξ 0 is given by:
ξ 0 = v F π Δ
For typical LTS materials like aluminum, we have:
v F 10 6 m / s , Δ 0.18 meV
yielding:
ξ 0 1.05 × 10 34 · 10 6 π · 1.8 × 10 22 18.5 × 10 9 m = 18.5 nm
This distance spans hundreds of atomic spacings, reinforcing the fact that the so-called Cooper pair is not localized in space and hence not a physical molecule.

23.2. Real-Space Pairing Interpretation Is Metaphorical

Even though the BCS formalism refers to “pairs,” these are not particle pairs in the conventional sense. The pairing correlation function in real space is given by:
F ( r 1 , r 2 ) = Ψ BCS | ψ ( r 2 ) ψ ( r 1 ) | Ψ BCS
This function is nonlocal and extends over ξ 0 . It does not peak sharply at any single | r 1 r 2 | , as would be the case for a localized bound state. The BCS state is better viewed as a coherent deformation of the Fermi sea rather than a collection of discrete bound objects [5].

23.3. Momentum-Space Nature of Pairing

The BCS state minimizes energy by introducing a new quasiparticle vacuum where pairs of states ( k , k ) are coherently filled. This minimizes the Hamiltonian:
H = k , σ ϵ k c k , σ c k , σ k , k V k , k c k c k c k c k
The self-consistent gap equation derived from this is:
Δ k = k V k , k Δ k 2 E k tanh E k 2 k B T
which reflects pairing in k -space, not r -space.

23.4. Interpretational Implications

Given this structure, the common narrative of electrons forming “molecule-like” bound pairs is misleading. The BCS state does not consist of localized objects, but a collective, phase-coherent modification of the many-body wavefunction. Statements like “Cooper pairs flow without resistance” should be interpreted carefully: what flows is the condensate phase, not individual physical entities. The absence of real-space localization undermines metaphors often used in popular and even academic treatments.

23.5. Conclusion

The BCS ground state is a momentum-space construct, not a spatially bound object. Even in low-temperature superconductors where BCS theory matches experiments well, interpreting the superconducting state as made of real-space electron pairs is inaccurate. The pairing is an emergent feature of the quasiparticle vacuum, not a reflection of bound particle states. This suggests that the pairing narrative, while computationally effective, should be viewed as a metaphor rather than a literal description.

24. Gauge Symmetry Breaking vs. Pairing as the Primary Mechanism of Superconductivity

The standard BCS theory postulates that superconductivity originates from the formation of Cooper pairs, i.e., weakly bound electron pairs with opposite momentum and spin. However, in modern field-theoretic language, superconductivity is alternatively viewed as the spontaneous breaking of a global U ( 1 ) gauge symmetry. This perspective, particularly through the Ginzburg–Landau and Anderson-Higgs mechanisms, raises an important conceptual possibility: that Cooper pairing is not the primary cause but rather a low-energy emergent consequence of a more fundamental symmetry-breaking process.

24.1. Ginzburg–Landau Framework and U ( 1 ) Symmetry

In the Ginzburg–Landau (GL) theory, the complex order parameter ψ ( r ) = | ψ | e i θ ( r ) plays a central role. The GL free energy is:
F [ ψ ] = d 3 x α | ψ | 2 + β 2 | ψ | 4 + 1 2 m * i 2 e A ψ 2 + | B | 2 2 μ 0
Minimizing F [ ψ ] with respect to ψ yields a non-zero expectation value ψ 0 when α < 0 , signaling spontaneous breaking of the U ( 1 ) phase symmetry.
This framework does not demand a physical “pair” of electrons in real space. The magnitude | ψ | 2 corresponds to superfluid density, and the phase θ defines long-range coherence, which governs observable phenomena such as the Josephson effect and flux quantization.

24.2. Anderson-Higgs Mechanism in Condensed Matter

The Anderson mechanism explains how Goldstone modes arising from broken symmetry become “massive” due to coupling with the electromagnetic field [38]. Consider small fluctuations in the order parameter:
ψ ( r ) = | ψ 0 | + δ ρ ( r ) e i θ ( r )
Coupling to the electromagnetic potential A via minimal substitution modifies the kinetic term:
i 2 e A ψ 2 = | ψ 0 | 2 θ 2 e A 2
This term leads to a mass term for the vector potential:
L mass = 1 2 m A 2 A 2 , with m A 2 = 4 e 2 | ψ 0 | 2 2
indicating that electromagnetic waves become gapped inside the superconductor. The Meissner effect is thus a consequence of gauge symmetry breaking, not necessarily of Cooper pairing.

24.3. Alternative Models Without Real-Space Pairing

It is entirely plausible to construct models that exhibit U ( 1 ) symmetry breaking without invoking the specific pairing mechanism of BCS. Consider a bosonic condensate with long-range coherence. For instance, a hypothetical system of spin-singlet bosons formed through many-body entanglement or emergent phenomena could achieve ψ 0 without a pairing Hamiltonian.
Moreover, in the limit where the coherence length ξ , phase fluctuations dominate, and the concept of a localized Cooper pair loses meaning. This perspective aligns with topological superconductors and quantum phase liquid models, where pairing is emergent and not fundamental.

24.4. Critical Comparison with BCS

The BCS wavefunction:
| Ψ BCS = k u k + v k c k c k | 0
does not exhibit symmetry breaking explicitly, but its ground state does not conserve particle number. When rewritten using a number-conserving Bogoliubov transformation, the BCS state still reflects global phase coherence. Therefore, it is more accurate to say that phase coherence is the defining trait of the superconducting state, not the existence of discrete, real-space “pairs.”

24.5. Philosophical Implications

If gauge symmetry breaking is the true driver of superconductivity, then the entire narrative around pairing becomes secondary or possibly misleading. This shifts focus from microscopic interactions (e.g., phonons or spin fluctuations) to macroscopic order parameters and topology. The Higgs-like mass term for photons and the appearance of flux quantization are better understood as emergent properties of broken symmetry, not proof of a two-body bound state.

24.6. Conclusion

Viewing superconductivity through the lens of gauge symmetry breaking offers a more general and possibly more fundamental framework than Cooper pairing. The phase-coherent order parameter, Meissner effect, and Anderson-Higgs mass generation all support the interpretation that long-range coherence, rather than discrete electron pairing, defines the superconducting state. Future theories may benefit from decoupling the concept of superconductivity from the necessity of pair formation and instead anchor it in symmetry and topological considerations.

25. Superconductivity as Macroscopic Entanglement

The standard interpretation of superconductivity via the BCS framework emphasizes pairing of electrons with opposite momentum and spin to form a macroscopically occupied quantum ground state. However, an alternative viewpoint suggests that what truly distinguishes the superconducting state is not pairing per se, but the emergence of macroscopic quantum entanglement among many-body fermionic degrees of freedom. This interpretation hinges on the notion of phase coherence as a signature of collective quantum correlations, rather than localized two-particle binding.

25.1. BCS State as an Entangled Resource

The BCS ground state is given by
| Ψ BCS = k u k + v k c k c k | 0
This expression is an entangled state in Fock space, since it cannot be factorized into independent states of individual momentum modes. Each momentum mode is correlated with its time-reversed partner. When viewed in the occupation number basis, this state resembles a generalized Greenberger–Horne–Zeilinger (GHZ) state over the set of ( k , k ) pairs.
To quantify entanglement, one can compute the von Neumann entropy associated with tracing out a subset of momentum modes. Consider the reduced density matrix for a single pair ( k , k ) :
ρ k = Tr k | Ψ BCS Ψ BCS |
The corresponding entropy is
S k = Tr ρ k log ρ k = | u k | 2 log | u k | 2 | v k | 2 log | v k | 2
Maximal entanglement occurs for | u k | 2 = | v k | 2 = 1 2 , which is near the Fermi surface, where the gap opens.

25.2. Mutual Information and Global Coherence

Another route is to compute the mutual information between spatial regions of the superconductor. Let A and B denote subsystems, then mutual information is:
I ( A : B ) = S A + S B S A B
For the BCS state, long-range coherence implies significant mutual information even between spatially separated regions, unlike in a Fermi gas. This supports the idea that superconductivity embodies a nonlocal quantum resource.

25.3. Comparison with Other Entangled States

The BCS state shares similarities with cluster states used in measurement-based quantum computing. In both cases, the global state is entangled, and measurements on one part can affect outcomes in another. However, BCS states are Gaussian fermionic states, whereas cluster states are typically spin-based and non-Gaussian.
According to [43], the BCS state contains multipartite entanglement over the entire Fermi sea. Vedral argues [42] that superconductivity is an example of “useful entanglement” that can be harnessed for quantum technological applications. Amico et al. [44] further review how macroscopic entanglement arises in various condensed matter systems, with superconductors as a prototypical case.

25.4. Interpretational Shift

From this perspective, the energy gap Δ and critical temperature T c are secondary to the emergence of entanglement. What defines the superconducting state is the transition from a separable Fermi gas to a highly entangled condensate with a coherent phase. This transition is captured not merely by the onset of pairing but by a sudden global correlation structure.
Thus, the BCS wavefunction, while derived from microscopic interactions, should be interpreted as a many-body entangled state rather than a collection of electron pairs.

25.5. Conclusion

The superconducting condensate is best understood as a macroscopic quantum-entangled state. This interpretation emphasizes long-range coherence and nonlocal correlations over real-space pairing. It aligns with experimental phenomena such as phase stiffness, flux quantization, and the robustness of the Josephson effect. By focusing on entanglement as the primary order parameter, we may uncover a deeper understanding of superconductivity, one that transcends the Cooper pair metaphor and leads to novel theoretical and experimental insights.

26. Revisiting the Role of Fermi Liquid Theory in Low-Temperature Superconductors

The Bardeen-Cooper-Schrieffer (BCS) theory of superconductivity assumes that the normal metallic state from which superconductivity emerges is a Fermi liquid. This foundational assumption implies that well-defined quasiparticles exist with a one-to-one correspondence to non-interacting electrons, though renormalized by interactions. However, empirical evidence and theoretical developments suggest that deviations from Fermi liquid behavior may already exist in low-temperature superconductors (LTS).

26.1. Mass Renormalization and the Gap Equation

Within Fermi liquid theory, the effective mass m * of the quasiparticles differs from the bare electron mass m due to interactions. The ratio m * / m is critical in determining the density of states at the Fermi level N ( 0 ) , which directly enters the BCS gap equation:
Δ = 2 ω D exp 1 N ( 0 ) V
where ω D is the Debye frequency and V is the pairing interaction. Since
N ( 0 ) = m * k F π 2 3
a small change in m * can exponentially affect Δ . For aluminum, with m * / m 1.05 , and for lead, m * / m 1.3 from quantum oscillation measurements [45], this variation contributes to observed gaps between Δ Al 0.18 meV and Δ Pb 1.3 meV .

26.2. Quantum Oscillations and Quasiparticle Lifetimes

de Haas–van Alphen (dHvA) and Shubnikov–de Haas (SdH) oscillations provide direct measurements of the Fermi surface topology and quasiparticle effective masses. The amplitude of these oscillations is described by the Lifshitz–Kosevich formula:
A ( T , B ) X sinh X , X = 2 π 2 k B T m * e B
Precise fits of this expression allow extraction of m * , with experimental uncertainties often below 5%. These studies for Pb and Sn reveal slight but non-negligible enhancements over m, supporting renormalization effects due to electron-phonon interactions [46].

26.3. Marginal Fermi Liquid Considerations

Although Landau Fermi liquid theory posits well-defined quasiparticles with lifetime τ ϵ 2 , the marginal Fermi liquid (MFL) model proposed for cuprates suggests instead τ 1 ϵ [47]. While MFL has not been widely applied to LTS, recent high-resolution photoemission studies indicate possible deviations even in conventional metals near T c , such as broadened linewidths inconsistent with Fermi liquid expectations.

26.4. Implications for the BCS Assumptions

If the underlying metallic state of LTS deviates from a pure Fermi liquid, then the BCS foundation becomes less solid. The pairing interaction might then be more sensitive to local density fluctuations or vertex corrections not captured in the simple mean-field approach. Additionally, the BCS gap equation, being logarithmically divergent, relies crucially on a well-defined Fermi surface. Any smearing of the Fermi edge by non-Fermi-liquid behavior alters the thermodynamics.

26.5. Conclusion

The Fermi liquid foundation of the BCS theory in low-temperature superconductors merits closer scrutiny. Empirical data shows that effective mass enhancements and possible deviations from Landau behavior exist, though subtle. These may affect the pairing mechanism, density of states, and even the nature of the superconducting transition. Future experimental precision in spectroscopic and transport probes will be essential to disentangle the role of quasiparticle renormalization.

27. Weak-Coupling Instability Argument Re-Examined

One of the foundational claims of the BCS theory is that an arbitrarily weak attractive interaction between electrons near the Fermi surface destabilizes the Fermi sea, resulting in the formation of Cooper pairs. This leads to the claim that any attractive potential, no matter how small, causes a superconducting instability. While this is mathematically true within the idealized BCS model, its physical relevance is less robust when realistic effects such as Coulomb repulsion, and screening is taken into account.

27.1. The Cooper Instability

The original derivation by Cooper showed that for a pair of electrons in a filled Fermi sea, an attractive interaction V < 0 leads to a bound state with negative energy, no matter how small the interaction strength is. Consider two electrons with opposite momentum near the Fermi surface, their Schrödinger equation in the simplified form is
E ψ ( k ) = 2 ϵ k ψ ( k ) + k V k k ψ ( k )
Assuming V k k = V constant within a narrow shell of width ω D , and zero otherwise, the solution yields a bound state energy
E = 2 ϵ F 2 ω D exp 1 N ( 0 ) V
with N ( 0 ) the density of states at the Fermi level. Hence, any V > 0 results in E < 2 ϵ F , suggesting an instability. However, this assumes no Coulomb repulsion and perfect screening.

27.2. Limitations from Coulomb Repulsion

The Coulomb interaction in real metals is long-range and repulsive. Screening mitigates its range, introducing a Thomas–Fermi screening wavevector:
q TF 2 = 4 π e 2 N ( 0 ) ϵ 0
Even with screening, residual repulsion exists. The net interaction becomes
V eff ( q ) = V ph ( q ) 1 + V C ( q ) Π ( q )
where V ph is the attractive phonon-mediated interaction, V C is the Coulomb repulsion, and Π is the polarization function. For V ph to dominate, it must exceed the residual V C , which is non-trivial at low temperatures.

27.3. Failure of Weak-Coupling in Realistic Scenarios

As shown by Leggett [49], real materials often exhibit situations where weak coupling does not generate pairing due to competing interactions. For example, the net coupling constant λ μ * , where λ is the electron-phonon strength and μ * the Coulomb pseudopotential, must satisfy
λ μ * > 0
But μ * is not infinitesimal and is typically in the range 0.1 0.2 , which sets a threshold for λ . If λ < μ * , the Cooper instability does not arise. Hence, BCS’s universal instability claim does not hold in practice.

27.4. Kohn–Luttinger Mechanism

An alternative mechanism proposed by Kohn and Luttinger [48] showed that even in purely repulsive systems, superconductivity can emerge via higher-order corrections. The key idea is that effective attractive interactions appear in high angular momentum channels due to Friedel oscillations in the screened Coulomb potential. This gives an effective pairing interaction in the l-th channel as:
V l eff 1 l 4
This allows for superconductivity at low but finite temperatures, even when V > 0 everywhere. However, the resulting T c is extremely low, often negligible for practical materials, making it more of a theoretical curiosity than a robust alternative.

27.5. Conclusion

The assumption of a universal Cooper instability underpins much of the BCS formalism but fails to account for realistic effects such as Coulomb repulsion and screening. While the BCS formalism remains powerful for explaining many features of superconductivity, its reliance on an idealized weak-coupling argument is a conceptual weakness. More sophisticated models like those incorporating the Kohn–Luttinger mechanism or incorporating retardation effects beyond Migdal–Eliashberg theory may offer more insight.

28. Possible Hydrodynamic Analogs Without Pairing

The standard BCS theory interprets superconductivity as a consequence of electron pairing and condensation into a macroscopically coherent state. However, one can alternatively approach superconductivity from a hydrodynamic perspective, akin to the description of superfluid 4He or 1D Tonks–Girardeau gases. This analogy is particularly intriguing in light of the lack of direct real-space detection of Cooper pairs and the formal resemblance between superfluid and superconducting states.

28.1. Quantum Hydrodynamic Description

In hydrodynamic theory, the superfluid component is described by a complex order parameter Ψ ( r , t ) with amplitude and phase. The density and current are then given by
n s ( r , t ) = | Ψ ( r , t ) | 2 , j s = m Im Ψ * Ψ
The dynamics of the superfluid can be captured via the quantum Euler equations. The continuity equation is
n s t + · ( n s v s ) = 0
and the Euler-like momentum equation, incorporating the quantum pressure term, is
m v s t + v s · v s = μ + 2 2 m 2 n s n s
These equations allow for vortex solutions, quantized in units of circulation
v s · d l = 2 π m n
where n is an integer and m is the effective mass of the fluid particle. If the constituent is a single electron (not a pair), the circulation quantum becomes / e , not / 2 e , offering a testable prediction [50].

28.2. Superfluid Density and Phase Rigidity

The notion of superfluid density n s corresponds in superconductors to the density of phase-coherent charge carriers. The London equations can be reinterpreted in this context:
j s = n s e 2 m A
Combining with Maxwell’s equations yields the London penetration depth:
λ L = m μ 0 n s e 2
Here, the identification of n s as independent from pairing becomes possible. That is, one can envision n s as a coherent fluid density, similar to the condensate fraction in 4He, without requiring a microscopic pair structure.

28.3. Analogy to Tonks–Girardeau Gases

In one-dimensional systems of strongly repulsive bosons, the Tonks–Girardeau gas forms a quantum degenerate state mimicking fermionic behavior. The bosonic wavefunction satisfies
Ψ B ( x 1 , , x N ) = 1 i < j N sgn ( x i x j ) Ψ F ( x 1 , , x N )
where Ψ F is a Slater determinant. The system behaves hydrodynamically but without true bosonic condensation. This analogy suggests a model where quantum statistics and strong correlations, not pairing, yield coherent behavior.

28.4. Vortex Quantization Without Pairing

In superconductors, quantized vortices are typically attributed to the phase winding of the Cooper pair condensate. However, quantized circulation does not intrinsically require pairs. If the order parameter is a single-electron phase field, then flux quantization should be in units of h / e , rather than h / 2 e . Some exotic systems, such as half-quantum vortices in Sr2RuO4, hint at the possibility of such scenarios [51].

28.5. Conclusion

The hydrodynamic analogy offers a route to understanding superconductivity without invoking real-space Cooper pairing. By focusing on macroscopic phase coherence, fluid rigidity, and quantum entanglement, one can derive many phenomenological features of superconductors. Further experimental tests, such as searching for h / e flux quantization or modified superfluid density behavior, could distinguish between pairing-based and hydrodynamic origins of superconductivity.

29. Possible Hydrodynamic Analogs Without Pairing

The standard BCS theory interprets superconductivity as a consequence of electron pairing and condensation into a macroscopically coherent state. However, one can alternatively approach superconductivity from a hydrodynamic perspective, akin to the description of superfluid 4He or 1D Tonks–Girardeau gases. This analogy is particularly intriguing in light of the lack of direct real-space detection of Cooper pairs and the formal resemblance between superfluid and superconducting states.

29.1. Quantum Hydrodynamic Description

In hydrodynamic theory, the superfluid component is described by a complex order parameter Ψ ( r , t ) with amplitude and phase. The density and current are then given by
n s ( r , t ) = | Ψ ( r , t ) | 2 , j s = m Im Ψ * Ψ
The dynamics of the superfluid can be captured via the quantum Euler equations. The continuity equation is
n s t + · ( n s v s ) = 0
and the Euler-like momentum equation, incorporating the quantum pressure term, is
m v s t + v s · v s = μ + 2 2 m 2 n s n s
These equations allow for vortex solutions, quantized in units of circulation
v s · d l = 2 π m n
where n is an integer and m is the effective mass of the fluid particle. If the constituent is a single electron (not a pair), the circulation quantum becomes / e , not / 2 e , offering a testable prediction [50].

29.2. Superfluid Density and Phase Rigidity

The notion of superfluid density n s corresponds in superconductors to the density of phase-coherent charge carriers. The London equations can be reinterpreted in this context:
j s = n s e 2 m A
Combining with Maxwell’s equations yields the London penetration depth:
λ L = m μ 0 n s e 2
Here, the identification of n s as independent from pairing becomes possible. That is, one can envision n s as a coherent fluid density, similar to the condensate fraction in 4He, without requiring a microscopic pair structure.

29.3. Analogy to Tonks–Girardeau Gases

In one-dimensional systems of strongly repulsive bosons, the Tonks–Girardeau gas forms a quantum degenerate state mimicking fermionic behavior. The bosonic wavefunction satisfies
Ψ B ( x 1 , , x N ) = 1 i < j N sgn ( x i x j ) Ψ F ( x 1 , , x N )
where Ψ F is a Slater determinant. The system behaves hydrodynamically but without true bosonic condensation. This analogy suggests a model where quantum statistics and strong correlations, not pairing, yield coherent behavior.

29.4. Vortex Quantization Without Pairing

In superconductors, quantized vortices are typically attributed to the phase winding of the Cooper pair condensate. However, quantized circulation does not intrinsically require pairs. If the order parameter is a single-electron phase field, then flux quantization should be in units of h / e , rather than h / 2 e . Some exotic systems, such as half-quantum vortices in Sr2RuO4, hint at the possibility of such scenarios [51].

29.5. Conclusion

The hydrodynamic analogy offers a route to understanding superconductivity without invoking real-space Cooper pairing. By focusing on macroscopic phase coherence, fluid rigidity, and quantum entanglement, one can derive many phenomenological features of superconductors. Further experimental tests, such as searching for h / e flux quantization or modified superfluid density behavior, could distinguish between pairing-based and hydrodynamic origins of superconductivity.

30. Non-Perturbative Many-Body Theories Beyond BCS

The Bardeen-Cooper-Schrieffer (BCS) theory remains the cornerstone of conventional superconductivity, but it is fundamentally a mean-field approach. It neglects strong correlations and fluctuation effects that might become significant even in low-temperature superconductors. The electron-phonon interaction is treated via an effective attractive potential, linearized around the Fermi surface, without accounting for retardation and quantum fluctuations in a fully non-perturbative way. More rigorous treatments require working directly with the many-body Hamiltonian:
H = k , σ ϵ k c k σ c k σ + q ω q b q b q + k , q , σ g q c k + q σ c k σ ( b q + b q )
where c k σ and b q are electron and phonon annihilation operators respectively, ϵ k is the electron dispersion, ω q the phonon energy, and g q the coupling constant. BCS theory treats this Hamiltonian under a static, mean-field approximation, assuming pairing occurs only near the Fermi surface and decouples phonons adiabatically.

30.1. Quantum Monte Carlo and Sign Problem

Quantum Monte Carlo (QMC) offers a path to exact treatment of many-body systems. However, in fermionic systems, especially with electron-phonon interactions, the fermion sign problem often renders simulations exponentially expensive. The path-integral formulation of QMC leads to an effective action
Z = D [ ψ ¯ , ψ , ϕ ] e S [ ψ ¯ , ψ , ϕ ]
where the action S is defined on imaginary time. Integration over phonons yields an effective non-local electron interaction. When computing observables, cancellation due to sign alternation causes the statistical error to grow exponentially with system size [52].

30.2. Diagrammatic Resummation

Another approach involves summing infinite subsets of Feynman diagrams — such as ladder or parquet approximations. For instance, the two-particle Green’s function G ( 2 ) can be built from the Bethe-Salpeter equation:
G ( 2 ) = G G + G G · Γ · G ( 2 )
where Γ is the two-particle irreducible vertex. BCS corresponds to summing only the particle-particle ladder diagrams. Including particle-hole channels leads to competing instabilities, such as spin or charge density waves [53].

30.3. Density Matrix Renormalization Group (DMRG)

DMRG is a powerful technique especially suited to one-dimensional systems. It constructs a variational wavefunction optimized over a reduced Hilbert space by retaining the most significant Schmidt eigenstates. The entanglement entropy is given by the von Neumann measure:
S = Tr ( ρ A ln ρ A )
where ρ A is the reduced density matrix of a subsystem. DMRG calculations on electron-phonon chains (e.g., Holstein or SSH models) show pairing without off-diagonal long-range order, questioning whether superconductivity must always involve pair condensation [54].

30.4. Beyond Mean-Field: Retardation and Vertex Corrections

Migdal’s theorem suggests vertex corrections are small due to the ratio ω D / E F 1 . However, in strong coupling or low-dimensional systems, this assumption fails. The full Eliashberg equations include frequency-dependent self-energy corrections:
Δ ( i ω n ) = π T m λ ( i ω n i ω m ) Δ ( i ω m ) ω m 2 + Δ 2 ( i ω m )
where λ is the spectral function of the electron-phonon interaction. Numerical solutions of these equations can differ from BCS predictions, especially in gap anisotropy and isotope effects [55].

30.5. Conclusion

While BCS theory captures much of the phenomenology, it should not be mistaken for a complete theory. More accurate treatments of the full electron-phonon many-body problem are limited by computational complexity and lack of analytical control. It remains possible that the real mechanism of superconductivity — even in LTS — involves physics beyond the scope of mean-field pairing.

31. Failure of BCS in 1D and 2D Systems: A Dimensional Critique

The Bardeen-Cooper-Schrieffer (BCS) theory assumes long-range phase coherence is spontaneously established once electron pairs condense into a ground state. However, in reduced dimensions — specifically one- and two-dimensional systems — such long-range order is fundamentally unstable against thermal and quantum fluctuations. This contradiction suggests that pairing, while present, may not be the true order parameter in these systems.

31.1. The Mermin-Wagner-Hohenberg Theorem

In dimensions d 2 , the Mermin-Wagner theorem asserts that continuous symmetries cannot be spontaneously broken at finite temperature in systems with short-range interactions. This prohibits the development of a long-range superconducting order parameter ψ ψ in 1D and 2D systems:
ψ ( r ) ψ ( r ) = 0 for T > 0 , d 2
This challenges the very foundation of BCS, which relies on a uniform condensate wavefunction across the entire system [56].

31.2. Berezinskii-Kosterlitz-Thouless (BKT) Transition

Despite the lack of true long-range order in 2D, superconductivity is still observed in thin films. This phenomenon is attributed to the BKT transition, a topological phase transition marked by the unbinding of vortex-antivortex pairs at the critical temperature T BKT . The superfluid stiffness ρ s jumps discontinuously at T BKT :
ρ s ( T BKT ) = 2 π T BKT
The transition is not characterized by a local order parameter but by topological defects and algebraically decaying correlations [57,58].

31.3. Quantum Fluctuations and 1D Luttinger Liquids

In strictly one-dimensional conductors, such as carbon nanotubes or organic chains, the low-energy excitations are described not by fermionic quasiparticles but by collective bosonic modes of charge and spin density waves. The Luttinger liquid theory provides the framework. The correlation function of a superconducting order parameter decays algebraically:
ψ ( x ) ψ ( 0 ) | x | 1 / K
where K is the Luttinger parameter. For K > 1 , superconducting fluctuations dominate over charge-density wave fluctuations, but no true long-range superconducting order is formed [59].

31.4. Quantum Criticality in 2D Superconductors

Thin-film superconductors such as InOx, TiN, and NbN display a quantum superconductor-insulator transition (QSIT) driven by magnetic field or disorder. Near the quantum critical point, the resistance scales as
R ( T ) = R c f T | B B c | z ν
where z and ν are critical exponents. Such transitions do not conform to BCS-like pair-breaking scenarios but rather to phase coherence suppression, suggesting a breakdown of the mean-field description [60].

31.5. Implications for BCS Theory

The persistence of pairing in 1D and 2D systems without global phase coherence implies that coherence — not just pairing — is the true driver of superconductivity. BCS theory, with its focus on momentum-space pairing and mean-field coherence, cannot capture this nuance. Real superconductivity in low-dimensional systems appears to be governed by topological or quantum fluctuation phenomena rather than simple pair condensation.

32. Optical Conductivity Spectra and Missing Sum Rule Weight

The optical conductivity σ 1 ( ω ) in superconductors exhibits a sharp delta function at ω = 0 due to the dissipationless current, and a gapped response at finite frequencies. The Ferrell-Glover-Tinkham (FGT) sum rule relates the total spectral weight in the normal and superconducting states. However, in many materials, especially strong-coupling superconductors, the transfer of spectral weight is not accounted for cleanly by conventional BCS theory, raising questions about its completeness.

32.1. Ferrell-Glover-Tinkham Sum Rule

The FGT sum rule expresses the conservation of spectral weight:
0 σ 1 N ( ω ) σ 1 S ( ω ) d ω = π n s e 2 2 m
where σ 1 N ( ω ) is the real part of the optical conductivity in the normal state, and σ 1 S ( ω ) in the superconducting state. The right-hand side represents the superfluid density n s . This implies that the area missing under σ 1 ( ω ) in the superconducting state appears as a delta-function at ω = 0 .

32.2. BCS Prediction and Experimental Deviations

In weak-coupling BCS theory, the conductivity is gapped below 2 Δ , where Δ is the superconducting gap. For ω < 2 Δ ,
σ 1 ( ω ) 0 for T T c , ω < 2 Δ
and above 2 Δ , the conductivity rises due to pair-breaking excitations. The total spectral weight lost in the finite- ω conductivity must match the delta-function contribution:
0 σ 1 S ( ω ) d ω + D = 0 σ 1 N ( ω ) d ω
where D is the weight of the delta function at ω = 0 .
However, in strong-coupling superconductors such as Pb or Nb, or in high- T c cuprates, this balance does not hold precisely. For example, Homes et al. [61] observed discrepancies in integrated spectral weight, suggesting additional channels of conductivity not captured by BCS.

32.3. Spectral Weight Integration

Assuming an ideal Drude model in the normal state:
σ 1 N ( ω ) = σ 0 1 + ω 2 τ 2
with σ 0 = n e 2 τ / m , the integrated weight is:
0 σ 1 N ( ω ) d ω = π 2 · n e 2 m
In the superconducting state, missing weight Δ W defines the strength of the delta peak:
Δ W = 0 [ σ 1 N ( ω ) σ 1 S ( ω ) ] d ω = π n s e 2 2 m
But experimentally, Δ W often falls short, especially in the underdoped cuprates or in materials with strong electron-boson coupling.

32.4. Strong-Coupling Effects

Allen [62] and others extended BCS to strong-coupling via Eliashberg theory. The conductivity includes vertex corrections and energy-dependent self-energies, modifying the FGT sum rule. The integrated optical spectral weight becomes sensitive to renormalization functions Z ( ω ) :
σ ( ω ) = i ω p 2 4 π · 1 ω + i / τ * ( ω )
with τ * ( ω ) derived from Eliashberg equations. Deviations from FGT can be interpreted as signs of inelastic scattering, pseudogap effects, or incoherent background not captured by the delta-function formalism.

32.5. Conclusion

The FGT sum rule provides a powerful test of BCS theory. Its violation in strong-coupling superconductors or in systems with pseudogap behavior points to deeper many-body effects beyond mean-field pairing. The redistribution of spectral weight and its interpretation remain an active area of research.

33. Topological Obstructions to BCS Condensation

The BCS theory of superconductivity assumes that a uniform, isotropic gap can form on a Fermi surface, commonly associated with s-wave symmetry. However, advances in topological band theory have shown that certain Fermi surface geometries or topological invariants may preclude such condensation due to symmetry or topological obstructions. In this section, we explore how spin-orbit coupling, Berry curvature, and band topology can inhibit or reshape the nature of superconducting order.

33.1. Berry Phase and Fermi Surface Geometry

The presence of a non-zero Berry phase across the Fermi surface is a fundamental topological obstruction. In systems with significant Berry curvature Ω ( k ) , the phase accumulated by electrons around closed orbits influences the pairing channel. The Berry curvature is defined as:
Ω ( k ) = k × A ( k ) , A ( k ) = i u k | k | u k
Here, | u k is the periodic part of the Bloch wavefunction. A nontrivial Berry phase γ = C A ( k ) · d k can obstruct the construction of a globally defined s-wave pair potential, as it alters the transformation properties of the gap under crystal symmetries.

33.2. Spin-Orbit Coupling and Helical States

Strong spin-orbit coupling (SOC) entangles spin and momentum, leading to helical Fermi surfaces where each momentum state is paired with a spin-polarized partner. The Rashba Hamiltonian is a canonical example:
H R = p 2 2 m + α R ( σ x k y σ y k x )
Diagonalization yields two split bands:
E ± ( k ) = 2 k 2 2 m ± α R | k |
Conventional BCS pairing assumes ( k , ) pairs with ( k , ) , but in spin-split bands, time-reversed partners are no longer degenerate, complicating s-wave pairing. The system instead may favor triplet or odd-parity pairing states [63].

33.3. Topological Invariants and Pairing Symmetry

Topological insulators and semimetals possess band structures characterized by Z 2 invariants or Chern numbers. These invariants constrain the possible superconducting states. In the presence of a nontrivial Z 2 index, the bulk cannot support a conventional s-wave gap without breaking time-reversal symmetry or closing the gap.
Moreover, nodal structures in unconventional superconductors (e.g., d-wave or p-wave) can be protected by topological invariants. The winding number ν associated with the gap function Δ ( k ) on a closed Fermi surface loop C is:
ν = 1 2 π i C d k · k log Δ ( k )
A non-zero winding number implies nodes or sign changes in Δ ( k ) , precluding a simple s-wave pairing.

33.4. Physical Consequences in LTS

Even in low-temperature superconductors (LTS) such as Pb or Al, recent experimental studies suggest non-negligible SOC or Berry curvature due to the heavy atomic constituents or surface effects. In heterostructures or ultrathin films, Rashba SOC can become prominent, modifying the pairing landscape. In these cases, deviations from ideal s-wave behavior, including admixtures of triplet components, have been observed [64].
Moreover, certain materials may possess Dirac or Weyl points near the Fermi level, introducing monopoles of Berry curvature that act as topological obstructions to conventional pairing.

33.5. Conclusion

The BCS framework implicitly assumes a topologically trivial Fermi surface that permits isotropic s-wave condensation. However, in systems with nontrivial band topology, Berry curvature, or strong spin-orbit coupling, such condensation is obstructed or modified. These findings suggest that the success of BCS in LTS may rely on accidental topological triviality rather than a fundamental universality.

34. Reinterpreting the Energy Gap as a Many-Body Correlation Scale

The traditional view in BCS theory interprets the superconducting gap Δ as the binding energy of a Cooper pair. This interpretation assumes that Δ represents the energy needed to break a two-particle bound state. However, a deeper look into the many-body formalism of superconductivity suggests that Δ may be more accurately viewed as a many-body correlation scale characterizing collective phenomena, not a two-particle molecular object.
In BCS theory, the energy gap is derived from the gap equation,
1 = V k 1 2 E k , E k = ( ε k μ ) 2 + Δ 2
where V is the effective interaction, ε k is the electron dispersion, and μ the chemical potential. The energy gap appears here as a variational parameter in the minimization of the BCS energy functional, not as a two-particle binding energy from a Bethe–Salpeter equation. This already hints at the collective nature of Δ .
From a correlation function perspective, the anomalous Green’s function (pair propagator),
F ( k , τ ) = T τ c k ( τ ) c k ( 0 )
has a long-time decay governed by Δ , indicating that Δ sets the temporal correlation length of Cooper-like correlations. Furthermore, the spectral gap in the single-particle Green’s function,
G ( k , ω ) = u k 2 ω E k + i η + v k 2 ω + E k i η
also indicates a suppression of single-particle excitations below energy Δ , but this does not imply a real-space bound state.
An alternative framework to interpret Δ is through the entanglement spectrum of the BCS ground state. The reduced density matrix for a spatial subregion A shows an entanglement gap which correlates with Δ [65]. This suggests that Δ quantifies the many-body entanglement structure of the superconducting state. The von Neumann entropy,
S A = Tr [ ρ A log ρ A ]
exhibits scaling behavior that reflects the coherence scale set by Δ .
In systems exhibiting conformal invariance or described by conformal field theory (CFT), a similar gap arises in the operator scaling dimensions, which is associated with finite-size correlation energies [66]. This analogy reinforces the idea that Δ is a collective excitation gap emerging from the system’s correlated nature.
Strongly correlated electron systems also suggest reinterpreting Δ . For instance, in the Hubbard model with a nonzero pairing amplitude but suppressed long-range order, Δ may survive without a true superconducting phase, emphasizing its role as a measure of local pairing correlations rather than global coherence.
Thus, it is more accurate to view Δ not as the binding energy of a discrete two-electron object, but rather as a scale arising from the coherent many-body entanglement of the entire electron fluid. This re-framing of Δ aligns superconductivity conceptually with quantum critical systems, where energy gaps signify collective order rather than particle-level interactions.

35. Conclusion

While BCS theory and the Cooper pair concept explain many features of traditional superconductors, the theory’s extension to complex and strongly correlated systems reveals multiple inconsistencies. The mismatch in energy scales, unrealistic assumptions about coherence lengths, failure in high- T c materials, and the emergence of alternative mechanisms necessitate a re-evaluation of the fundamental assumptions of superconductivity. It is increasingly evident that the Cooper pair mechanism may be a special case of a broader, yet undiscovered, quantum condensation phenomenon.
Despite the profound success of the BCS theory in explaining numerous experimental features of conventional superconductors, a growing body of conceptual and empirical inconsistencies challenge its completeness and foundational interpretation. These concerns are not limited to exotic or high-temperature systems, but emerge even in canonical low-temperature superconductors such as aluminum, tin, and lead.
A central critique is the lack of direct evidence for real-space Cooper pairs. The BCS wavefunction, which operates in momentum space, describes a condensate of entangled states rather than localized bound objects. The coherence length ξ 0 in low-temperature superconductors far exceeds interatomic spacings, implying a delocalization incompatible with the naive image of “paired electrons.” Further, the fragility of the superconducting state — vulnerable to minuscule perturbations — contradicts the notion of an energetically robust paired ground state.
Moreover, experimental observations such as the small drift velocity of electrons, in stark contrast to the nearly instantaneous propagation of superconducting currents, highlight the dominance of collective quantum phase coherence rather than quasiparticle dynamics. The discrepancy between energy scales, with Δ E F , adds to the concern that the condensation mechanism is not deeply rooted in a stable two-body interaction, but perhaps emerges from deeper many-body correlations, symmetry breaking, or topological constraints.
Alternative frameworks — including topological superconductivity, hydrodynamic analogs, macroscopic entanglement, and gauge-theoretic interpretations — offer competing narratives that may explain superconductivity without invoking Cooper pairs as physical entities. For instance, spontaneous symmetry breaking of the global U ( 1 ) gauge symmetry may be more foundational than pairing, with the latter being a consequence rather than a cause. Similarly, reinterpretation of the superconducting gap Δ as a many-body correlation scale rather than a pair-binding energy leads to an enriched conceptual landscape.
Future progress will depend not only on further experimental refinement — such as precise measurements of optical conductivity sum rules, vortex dynamics in low-dimensional systems, or entanglement entropy in superconducting phases — but also on a broader philosophical openness to questioning entrenched models. A more nuanced theory may require non-perturbative many-body approaches beyond mean-field theory, accounting for the complex topology and quantum coherence inherent in condensed matter systems.
In conclusion, while BCS theory has had remarkable historical success, it may be best viewed as a phenomenological model with limited explanatory scope. A truly foundational theory of superconductivity — especially one applicable to all regimes — may require rethinking the role of Cooper pairs altogether, shifting the narrative toward collective quantum phenomena that transcend simple pairing pictures.

References

  1. Cooper, L. N. Bound electron pairs in a degenerate Fermi gas. Physical Review 1956, 104, 1189. [Google Scholar] [CrossRef]
  2. Franck, J. P. (1994). Experimental studies of the isotope effect in high temperature superconductors. In Physical Properties of High Temperature Superconductors IV, World Scientific.
  3. Anderson, P. W. The resonating valence bond state in La2CuO4 and superconductivity. Science 1987, 235, 235–1198. [Google Scholar] [CrossRef] [PubMed]
  4. Gunnarsson, O., Calandra. Colloquium: Electron-phonon interaction in conventional and unconventional superconductors. Rev. Mod. Phys. 2008, 75, 1085. [Google Scholar] [CrossRef]
  5. Schrieffer, J. R. (1964). Theory of Superconductivity, W. A. Benjamin, Inc.
  6. Shapiro, S. M. , Shirane, G., & Noda, Y. Inelastic neutron scattering study of phonons in lead and niobium. Phys. Rev. B 1975, 12, 489. [Google Scholar]
  7. Anderson, P. W. Theory of dirty superconductors. Journal of Physics and Chemistry of Solids 1959, 11, 26–30. [Google Scholar] [CrossRef]
  8. Giaever, I. Energy gap in superconductors measured by electron tunneling. Physical Review Letters 1960, 5, 147. [Google Scholar] [CrossRef]
  9. de Gennes, P. G. (1966). Superconductivity of Metals and Alloys, Addison-Wesley.
  10. Anderson, P. W. Plasmons, gauge invariance, and mass. Physical Review 1963, 130, 439. [Google Scholar] [CrossRef]
  11. Randeria, M. (1995). Crossover from BCS theory to Bose-Einstein condensation. In Bose-Einstein Condensation, Cambridge University Press.
  12. Damascelli, A. , Hussain, Z., & Shen, Z. X. Angle-resolved photoemission studies of the cuprate superconductors. Rev. Mod. Phys. 2003, 75, 473. [Google Scholar]
  13. Zhao, G. M. , Hunt, M. B., Keller, H., & Müller, K. A. Evidence for polaronic superconductivity in La2-xSrxCuO4 from oxygen-isotope effect studies. Nature 1997, 385, 236–239. [Google Scholar]
  14. Tranquada, J. M. , Sternlieb, B. J., Axe, J. D., Nakamura, Y., & Uchida, S. (1995). Evidence for stripe correlations of spins and holes in copper oxide superconductors. ( 375(6532), 561–563.
  15. Tsuei, C. C. , & Kirtley, J. R. Pairing symmetry in cuprate superconductors. Rev. Mod. Phys. 2000, 72(4), 969. [Google Scholar]
  16. Mazin, I. I. , Singh, D. J., Johannes, M. D., & Du, M. H. Unconventional superconductivity with a sign reversal in the order parameter of LaFeAsO1-xFx. Phys. Rev. Lett. 2008, 101, 057003. [Google Scholar] [PubMed]
  17. Pan, S. H. , O’Neal, J. P., Badzey, R. L., Chamon, C., Ding, H., Engelbrecht, J. R., ... & Davis, J. C. Microscopic electronic inhomogeneity in the high-Tc superconductor Bi2Sr2CaCu2O8+δ. Nature 2001, 413, 282–285. [Google Scholar] [PubMed]
  18. Monthoux, P. , Pines, D., & Lonzarich, G. G. Superconductivity without phonons. Nature 2007, 450, 1177–1183. [Google Scholar] [PubMed]
  19. Van Harlingen, D. J. Phase-sensitive tests of the symmetry of the pairing state in the high-temperature superconductors—Evidence for dx2-y2 symmetry. Rev. Mod. Phys. 1995, 67, 515. [Google Scholar] [CrossRef]
  20. Andreev, A. F. The thermal conductivity of the intermediate state in superconductors. Sov. Phys. JETP 1964, 19, 1228–1231. [Google Scholar]
  21. Yang, C. N. Concept of off-diagonal long-range order and the quantum phases of liquid He and of superconductors. Rev. Mod. Phys. 1962, 34, 694. [Google Scholar] [CrossRef]
  22. Zanardi, P. Quantum entanglement in fermionic lattices. Phys. Rev. A 2002, 65, 042101. [Google Scholar] [CrossRef]
  23. Sacépé, B. , Dubouchet, T., Chapelier, C., Sanquer, M., Ovadia, M., Shahar, D., & Dynes, R. C. Localization of preformed Cooper pairs in disordered superconductors. Nature Physics 2011, 7, 239–244. [Google Scholar]
  24. Tinkham, M. (1996). Introduction to Superconductivity, (2nd ed.). McGraw-Hill.
  25. Leggett, A. J. Superfluidity. Rev. Mod. Phys. 1999, 71, S318. [Google Scholar] [CrossRef]
  26. Behnia, K., & Jaccard, D. Second sound and the superfluid analogy of heat in solids. Journal of Physics: Condensed Matter 2016, 28, 015602.
  27. Scalapino, D. J. The case for dx2-y2 pairing in the cuprate superconductors. Physics Reports 1995, 250, 329–365. [Google Scholar] [CrossRef]
  28. Damascelli, A. , Hussain, Z., & Shen, Z. X. Angle-resolved photoemission studies of the cuprate superconductors. Rev. Mod. Phys. 2003, 75, 473. [Google Scholar]
  29. Hartnoll, S. A. , Herzog, C. P., & Horowitz, G. T. Building a holographic superconductor. Phys. Rev. Lett. 2008, 101, 031601. [Google Scholar]
  30. Kitaev, A. Y. Unpaired Majorana fermions in quantum wires. Physics-Uspekhi 2001, 44, 131. [Google Scholar] [CrossRef]
  31. Reznik, D. , Pintschovius, L., Ito, M., Iikubo, S., Sato, M., Goka, H., ... & Fujita, M. (2006). Electron–phonon coupling reflecting dynamic charge inhomogeneity in copper oxide superconductors. Nature 2006, 440, 1170–1173. [Google Scholar]
  32. Kong, Y. , Dolgov, O. V., Jepsen, O., & Andersen, O. K. (2001). Electron-phonon interaction in the normal and superconducting states of MgB2. Physical Review B.
  33. Boeri, L. , Dolgov, O. V., & Golubov, A. A. is LaFeAsO1-xFx an electron-phonon superconductor? Physical Review Letters 2008, 101, 026403. [Google Scholar]
  34. Tranquada, J. M. , Woo, H., Perring, T. G., Goka, H., Gu, G. D., Xu, G., ... & Fujita, M. Quantum magnetic excitations from stripes in copper oxide superconductors. Nature 2004, 429, 534–538. [Google Scholar]
  35. Read, N. , & Green, D. Paired states of fermions in two dimensions with breaking of parity and time-reversal symmetries and the fractional quantum Hall effect. Phys. Rev. B 2000, 61, 10267. [Google Scholar]
  36. Senthil, T. Symmetry-protected topological phases of quantum matter. Annual Review of Condensed Matter Physics 2015, 6, 299–324. [Google Scholar] [CrossRef]
  37. Parks, R. D. Superconductivity; Vol. 1 and 2. Marcel Dekker Inc, 1969. [Google Scholar]
  38. Anderson, P. W. Random-phase approximation in the theory of superconductivity. Phys. Rev. 1958, 112, 1900. [Google Scholar] [CrossRef]
  39. Tinkham, M. (2004). Introduction to Superconductivity, Dover Publications.
  40. Schrieffer, J. R. (1999). Theory of Superconductivity, Perseus Books.
  41. Weinberg, S. (1996). The Quantum Theory of Fields, Volume 2: Modern Applications, Cambridge University Press.
  42. Vedral, V. Entanglement in the second quantization formalism. Central European Journal of Physics 2003, 1, 289–306. [Google Scholar] [CrossRef]
  43. Zanardi, P. Quantum entanglement in fermionic lattices. Physical Review A 2002, 65, 042101. [Google Scholar] [CrossRef]
  44. Amico, L. , Fazio, R., Osterloh, A., & Vedral, V. Entanglement in many-body systems. Reviews of Modern Physics 2008, 80, 517. [Google Scholar]
  45. Shoenberg, D. (1984). Magnetic oscillations in metals, Cambridge University Press.
  46. McCollam, A. , Julian, S. R., Rourke, P. M. C., & Hussey, N. E. High-precision de Haas–van Alphen studies of heavy fermions. Physica B: Condensed Matter 2005, 1–5, 359–361. [Google Scholar]
  47. Varma, C. M. , Littlewood, P. B., Schmitt-Rink, S., Abrahams, E., & Ruckenstein, A. E. Phenomenology of the normal state of Cu-O high-temperature superconductors. Physical Review Letters 1989, 63, 1996. [Google Scholar]
  48. Kohn, W. , & Luttinger, J. M. New mechanism for superconductivity. Physical Review Letters 1965, 15(12), 524. [Google Scholar]
  49. Leggett, A. J. Superfluidity. Reviews of Modern Physics 1999, 71, S318. [Google Scholar] [CrossRef]
  50. Pitaevskii, L. P. , & Stringari, S. (2003). Bose-Einstein condensation, Oxford University Press.
  51. Kirtley, J. R. Upper limit on spontaneous supercurrents in Sr2RuO4. Physical Review B 2007, 76(1), 014526. [Google Scholar] [CrossRef]
  52. Loh, E. Y. , Gubernatis, J. E., Scalettar, R. T., White, S. R., Scalapino, D. J., & Sugar, R. L. Sign problem in the numerical simulation of many-electron systems. Physical Review B 1990, 41, 9301. [Google Scholar]
  53. Metzner, W. , Salmhofer, M., Honerkamp, C., Meden, V., & Schönhammer, K. Functional renormalization group approach to correlated fermion systems. Reviews of Modern Physics 2012, 84(1), 299. [Google Scholar]
  54. Jeckelmann, E. , & White, S. R. Density-matrix renormalization-group study of the polaron problem in the Holstein model. Physical Review B 1999, 59, 738. [Google Scholar]
  55. Marsiglio, F. , Carbotte, J. P., & Blezius, J. Retardation effects in electron-phonon superconductivity. Physical Review B 1988, 37, 4965. [Google Scholar]
  56. Mermin, N. D. , & Wagner, H. Absence of ferromagnetism or antiferromagnetism in one-or two-dimensional isotropic Heisenberg models. Physical Review Letters 1966, 17, 1133. [Google Scholar]
  57. Kosterlitz, J. M. Ordering, metastability and phase transitions in two-dimensional systems. Journal of Physics C: Solid State Physic 1973, 6, 1181. [Google Scholar] [CrossRef]
  58. Berezinskii, V. L. Destruction of long-range order in one-dimensional and two-dimensional systems having a continuous symmetry group I. Classical systems. Soviet Journal of Experimental and Theoretical Physics 1971, 32, 493. [Google Scholar]
  59. Giamarchi, T. (2003). Quantum Physics in One Dimension, Oxford University Press.
  60. Sachdev, S. (2011). Quantum Phase Transitions, Cambridge University Press.
  61. Homes, C. C. , Dordevic, S. V., Strongin, M., Bonn, D. A., Liang, R., Hardy, W. N., ... & Timusk, T. A universal scaling relation in high-temperature superconductors. Nature 2004, 430, 539–541. [Google Scholar]
  62. Allen, P. B. Optical conductivity and the sum rule in the theory of superconductivity. Physical Review B 1971, 3, 305. [Google Scholar] [CrossRef]
  63. Fu, L. , & Kane, C. L. Superconducting proximity effect and Majorana fermions at the surface of a topological insulator. Physical Review Letters 2008, 100(9), 096407. [Google Scholar]
  64. Sato, M.; Ando, Y. Topological superconductors: a review. Reports on Progress in Physics 2017, 80, 076501. [Google Scholar] [CrossRef]
  65. L. Amico, R. L. Amico, R. Fazio, A. Osterloh, and V. Vedral, “Entanglement in many-body systems”. Rev. Mod. Phys. 2008, 80, 517. [Google Scholar]
  66. P. Calabrese and J. Cardy, “Entanglement entropy and quantum field theory”. J. Stat. Mech. 2004, P06002.
  67. V. Vedral, Entanglement in the second quantization formalism. Cent. Eur. J. Phys. 2003, 1, 289–306.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.
Copyright: This open access article is published under a Creative Commons CC BY 4.0 license, which permit the free download, distribution, and reuse, provided that the author and preprint are cited in any reuse.
Prerpints.org logo

Preprints.org is a free preprint server supported by MDPI in Basel, Switzerland.

Subscribe

Disclaimer

Terms of Use

Privacy Policy

Privacy Settings

© 2026 MDPI (Basel, Switzerland) unless otherwise stated