1 Introduction

Over recent decades, with the great advances in cooling technologies, much attention was focused on investigating the behaviors of Cooper pairs near zero temperature. Among all physical quantities, the zero-temperature superfluid phase stiffness ρs(0) is a central parameter for describing zero-temperature Cooper pairs, since it can be exactly obtained by measuring magnetic penetration depths of superconducting materials. For copper oxide materials, there has been much interest for seeking the potential correlations between the transition temperature Tc and ρs(0). The earliest pattern was referred to as the Uemura relation [12] Tcρs(0), which works reasonably well for the underdoped materials. Later, a more universal relation, the Homes’ law [3,4,5,6] Tcρs(0)/σdc was found to hold regardless of underdoped, optimally doped, and overdoped materials, where σdc denotes the dc conductivity measured at approximately Tc. Theoretically, Homes’ law has been well-known as a mean-field result of the dirty-limit BCS theory [4, 78]. Despite these successes, some scholars questioned the validity of Homes’ law in highly underdoped and overdoped sides. For example, the relation between Tc and ρs(0) was found to be sub-linear in highly underdoped materials [9,10,11,12]. Recently, by investigating the overdoped side of the single-crystal La2 − xSrxCuO4 films, Bozovic et al. observed a two-class scaling law [13]:

$$ \left\{\begin{array}{cc}{T}_c=\alpha \cdotp {\rho}_s(0)+{T}_0,& {T}_c\ge {T}_M\\ {}{T}_c=\gamma \cdotp \sqrt{\rho_s(0)},& {T}_c\le {T}_Q\end{array}\right., $$
(1)

where TM ≈ 12 K, TQ ≈ 15 K, α = 0.37 ± 0.02, T0 = (7.0 ± 0.1) K, and γ = (4.2 ± 0.5) K1/2. The difference between TM and TQ implies that the two-class scaling law (1) is non-smoothly linked by linear and parabolic parts.

Equation (1) indicates that a parabolic scaling emerges in the highly overdoped side [13]. Since the two-class scaling law (1) differs significantly from Homes’ law, Bozovic et al. concluded that their experimental findings are incompatible with the mean-field description [13,14,15]. The linear part in Eq. (1) can be derived by using the dirty-limit BCS theory [4, 78] and therefore is a mean-field result; however, the parabolic part may hint potential new physics [13]. As a possible evidence, Bozovic et al. have observed that with increased doping (Tc → 0), La2 − xSrxCuO4 becomes more metallic, and increased doping induces a quantum phase transition from a superconductor to a normal metal [13,14,15]. This observation indicates that when Tc → 0, quantum fluctuations may play an important role for inducing the parabolic scaling in Eq. (1). In this paper, we propose a quantum partition function for describing quantum critical behaviors of zero-temperature Cooper pairs. Based on such a quantum partition function, we will exactly reproduce the two-class scaling law (1). Here, we adopt the natural units = c = kB = 1, where denotes the reduced Planck constant, c is the light speed, and kB is the Boltzmann constant.

2 Quantum partition function for zero-temperature Cooper pairs

The free energy density of zero-temperature Cooper pairs can be generally written as [16]:

$$ \mathcal{L}=\sigma \cdotp {\left|{\partial}_{\tau}\phi \left(\boldsymbol{q},\tau \right)\right|}^2+\eta \cdotp {\left|\mathbf{\nabla}\phi \left(\boldsymbol{q},\tau \right)\right|}^2+{\lambda}_2\cdotp {\left|\phi \left(\boldsymbol{q},\tau \right)\right|}^2+{\lambda}_4\cdotp {\left|\phi \left(\boldsymbol{q},\tau \right)\right|}^4, $$
(2)

where ϕ(q, τ) denotes the order parameter of zero-temperature Cooper pairs and it is a function of space q and imaginary time τ. Here, \( \tau \in \left[0,\frac{1}{T}\right] \) with the temperature T being 0. σ, η, λ2, and λ4 are phenomenological parameters [16].

If one denotes the zero-temperature superfluid phase stiffness by |ϕ(q, τ)|2, then, by applying Gor’kov’s Green function method [8] into the BCS theory at T = 0 and Tc ≈ 0, one can obtain [17]:

$$ \eta =1, $$
(3)
$$ {\lambda}_2={\lambda}_2\left({T}_c\right)=-\frac{24{\pi}^2{m}_e}{7\zeta (3)\cdotp {\varepsilon}_F}{T}_c^2, $$
(4)
$$ {\lambda}_4={\lambda}_4\left({T}_c,{\rho}_s(0)\right)=\frac{12{\pi}^2{m}_e}{7\zeta (3)\cdotp {\varepsilon}_F}\cdotp \frac{T_c^2}{\rho_s(0)}, $$
(5)

where \( {\rho}_s(0)=\frac{n_s(0)}{4{m}_e} \) and ns(0) denote zero-temperature superfluid phase stiffness [13] and zero-temperature superfluid density when materials are homogenous, ζ(x) is the Riemann zeta function, εF is the Fermi energy, and me is the rest mass of an electron. The derivation for Eqs. (3)–(5) can be found in Appendix 1, where we have clarified why Gor’kov’s method holds at T = 0.

Equations (3)–(5) are derived by using the BCS theory, which assumes that quantum fluctuations on all size scales are averaged out. Based on such an assumption of the mean-field, ns(0) is equal to the total number density of electrons in the normal state [8] and hence can be regarded as a constant. This is the standard explanation of the BCS theory. However, later we will observe that ns(0) changes with Tc as long as quantum fluctuations cannot be averaged out.

Due to Eqs (3), (4), and (5), σ is the unique phenomenological parameter in Eq. (2). In this paper, we order σ = 1 so that the free energy density (2) yields an exact relativistic form:

$$ \mathcal{L}\left({T}_c\right)={\left|{\partial}_{\tau}\phi \left(\boldsymbol{q},\tau \right)\right|}^2+{\left|\mathbf{\nabla}\phi \left(\boldsymbol{q},\tau \right)\right|}^2+{\lambda}_2\left({T}_c\right)\cdotp {\left|\phi \left(\boldsymbol{q},\tau \right)\right|}^2+{\lambda}_4\left({T}_c,{\rho}_s(0)\right)\cdotp {\left|\phi \left(\boldsymbol{q},\tau \right)\right|}^4. $$
(6)

It is easy to observe that the transition temperature Tc in Eq. (6) plays the role of temperature T in the classical Landau-Ginzburg free energy. Later, we will show that Tc = 0 is a potential critical point. To guarantee the self-consistency of Eq. (6), we need to verify that |ϕ(q, τ)|2 is the zero-temperature superfluid phase stiffness. To this end, the free energy density (6) is varied to obtain the field equation of Cooper pairs:

$$ {\partial}_{\tau}^2\phi \left(\boldsymbol{q},\tau \right)+{\nabla}^2\phi \left(\boldsymbol{q},\tau \right)-{\lambda}_2\phi \left(\boldsymbol{q},\tau \right)-2{\lambda}_4\cdotp {\left|\phi \left(\boldsymbol{q},\tau \right)\right|}^2\phi \left(\boldsymbol{q},\tau \right)=0. $$
(7)

For homogenous superconductors, Eq. (7) yields |ϕ(q, τ)|2 = − λ2/2λ4 = ρs(0), where Eqs. (4) and (5) have been used. Because ρs(0) denotes the zero-temperature superfluid phase stiffness of homogenous materials, |ϕ(q, τ)|2 indeed denotes the zero-temperature superfluid phase stiffness. This verifies the self-consistency of the free energy density (6).

Using the free energy density (6), we propose a quantum partition function (QPF) for zero-temperature Cooper pairs as follows:

$$ Z\left({T}_c,J,{J}^{\ast}\right)=\int {\left[\mathcal{D}\phi {\left(\boldsymbol{q},\tau \right)}^{\ast}\right]}_{\Lambda}\int {\left[\mathcal{D}\phi \left(\boldsymbol{q},\tau \right)\right]}_{\Lambda}{e}^{-\int d\tau \int {d}^D\boldsymbol{q}\left[\mathcal{L}\left({T}_c\right)-J\left(\boldsymbol{q},\tau \right)\phi \left(\boldsymbol{q},\tau \right)-J{\left(\boldsymbol{q},\tau \right)}^{\ast}\phi {\left(\boldsymbol{q},\tau \right)}^{\ast}\right]}, $$
(8)

where J(q, τ) denotes the external field, Λ is the momentum cutoff, and D is the dimension of superconducting materials.

From a perspective of effective field theory, a quantum field theory should be defined fundamentally with a cutoff Λ [18,19,20]. For the crystal materials, a rigid renormalization theory can be defined on a cubic lattice of a lattice unit:

$$ a=\frac{1}{\Lambda}, $$
(9)

where a denotes the minimal lattice constant. The physical meaning of Eq. (9) is that quantum fluctuations with wavelengths less than 2πa can be averaged out [19]. Weinberg also pointed out that [21] in solid-state physics, there really is a cutoff, the lattice spacing a, which one must take seriously in dealing with phenomena at similar length scales.

Since the momentum cutoff Λ is determined by a, there is no longer any phenomenological parameter in the QPF (8). Therefore, the validity of the QPF (8) can be justified by the experimental investigation result (1).

3 Parabolic Scaling

We assume that quantum fluctuations with wavelengths larger than 2πa cannot be averaged out. By the theory of critical phenomena, this means that the coefficients λ2(Tc) and λ4(Tc, ρs(0)) in Eq. (6) should receive the contributions from quantum fluctuations on these size scales. To evaluate the contributions, by applying the renormalization group approach to the QPF (8), one can obtain the renormalization group equationsFootnote 1 [17]:

$$ \frac{d{\lambda}_2\left({T}_c\right)}{dlnb}={\lambda}_2\left({T}_c\right)\cdotp \left(2-4{\hat{\lambda}}_4\right)+O\left({\hat{\lambda}}_4^2\right), $$
(10)
$$ \frac{d{\hat{\lambda}}_4}{dlnb}=\left(3-D\right)\cdotp {\hat{\lambda}}_4-10{\hat{\lambda}}_4^2+O\left({\hat{\lambda}}_4^3\right), $$
(11)

where the quantum dynamical exponent z is equal to 1 and

$$ {\hat{\lambda}}_4={\lambda}_4\left({T}_c,{\rho}_s(0)\right)\cdotp \frac{{\left(\pi \right)}^{\frac{D}{2}}{\Lambda}^{D-3}}{2{\left(2\pi \right)}^D\Gamma \left(\frac{D}{2}\right)}. $$
(12)

By Eqs. (10)–(12), it is easy to get a nontrivial fixed point:

$$ \left\{\begin{array}{c}{\lambda}_2\left({T}_c\right)\approx 0\\ {}{\lambda}_4\left({T}_c,{\rho}_s(0)\right)\approx \frac{3-D}{10}\cdotp \frac{2{\left(2\pi \right)}^D\Gamma \left(\frac{D}{2}\right)}{{\left(\pi \right)}^{\frac{D}{2}}{\Lambda}^{D-3}}\end{array}\right.. $$
(13)

λ2(Tc) and λ4(Tc, ρs(0)) are defined by Tc and ρs(0) via Eqs. (4) and (5). Substituting Eqs. (4) and (5) into Eq. (13) yields:

$$ \left\{\begin{array}{c}{T}_c\approx 0\\ {}{T}_c\approx \gamma (D)\cdotp \sqrt{\rho_s(0)}\end{array}\right., $$
(14)

where

$$ \gamma (D)=\sqrt{\left(3-D\right)\cdotp {\Lambda}^{3-D}\cdotp \frac{7{\left(2\pi \right)}^D\Gamma \left(\frac{D}{2}\right)\zeta (3)\cdotp {\varepsilon}_F}{60{\left(\pi \right)}^{\frac{D}{2}+2}{m}_e}}. $$
(15)

If we denote Tc ≈ 0 by TcTQ(D), Eq. (14) can be written in the form:

$$ {T}_c=\gamma (D)\cdot \sqrt{\rho_s(0)}\kern0.5em \mathrm{for}\kern0.5em {T}_c\le {T}_Q(D), $$
(16)

where TQ(D) denotes a sufficiently low temperature. The physical meaning of Eq. (16) is that ρs(0) will change with Tc as long as TcTQ(D). Later, we will theoretically show TQ(2) ≤ γ(2)2 and TQ(3) ≤ 0.

The two-class scaling law (1) was found in the single-crystal La2 − xSrxCuO4 films (D = 2) around x = 0.25 [13]. Therefore, for D = 2, Eq. (16) reproduces the parabolic part in the two-class scaling law (1). To verify this, we show that γ(2) is in accordance with the existing experimental measure value. Plugging Eq. (9) into Eq. (15), one can obtain [21]:

$$ \gamma (2)=\sqrt{\frac{7\cdotp \zeta (3)\cdotp {\varepsilon}_F}{15\cdotp \pi \cdotp a\cdotp {m}_e}}. $$
(17)

For single-crystal La2 − xSrxCuO4 films, substituting the data a ≈ 3.8 × 10−10 m [13] and εF(x ≈ 0.2) ≈ 8.75 eV [22] into Eq. (17) yields [21]:

$$ \gamma (2)\approx 4.29\ {K}^{1/2}, $$
(18)

which exactly agrees with the experimental value (4.2 ± 0.5) K1/2 [13].

The high accordance between theoretical and experimental values thoroughly proves that the parabolic scaling in Eq. (1) is due to quantum fluctuations. From this meaning, the nontrivial fixed point (13) describes the quantum critical behaviors of zero-temperature Cooper pairs when TcTQ(D). However, we do not clarify the range of applicability of the nontrivial fixed point (13), i.e., the value of TQ(D). According to the renormalization group theory, the nontrivial fixed point (13) is valid if and only if quantum fluctuations cannot be averaged out. Therefore, to evaluate TQ(D), we need to find a criterion for identifying the validity of the mean-field approximation.

4 Quantum Ginzburg Number

For thermal fluctuations, there exists a clear criterion of the applicability of the mean-field theory, i.e., the classical Ginzburg number Gi [23,24,25], where the mean-field approximation is valid when Gi ≪ 1. To evaluate quantum fluctuations, we extend Gi to a quantum version. To this end, let us first define the correlation function of the order parameter ϕ(q, τ) as [16]:

$$ G\left(\boldsymbol{q}-{\boldsymbol{q}}^{\prime },\tau -{\tau}^{\prime}\right)=\left\langle \left[\phi \left(\boldsymbol{q},\tau \right)-\left\langle \phi \left(\boldsymbol{q},\tau \right)\right\rangle \right]\cdotp \left[\phi {\left({\boldsymbol{q}}^{\prime },{\tau}^{\prime}\right)}^{\ast }-\left\langle \phi {\left({\boldsymbol{q}}^{\prime },{\tau}^{\prime}\right)}^{\ast}\right\rangle \right]\right\rangle, $$
(19)

where the mean value of a physical variable A(q, τ) is defined by

$$ \left\langle A\left(\boldsymbol{q},\tau \right)\right\rangle =\frac{1}{Z\left({T}_c,J,{J}^{\ast}\right)}\int \mathcal{D}\phi {\left(\boldsymbol{q},\tau \right)}^{\ast}\int \mathcal{D}\phi \left(\boldsymbol{q},\tau \right){e}^{-\int d\tau \int {d}^D\boldsymbol{q}\left[\mathcal{L}\left({T}_c\right)-J\left(\boldsymbol{q},\tau \right)\phi \left(\boldsymbol{q},\tau \right)-J{\left(\boldsymbol{q},\tau \right)}^{\ast}\phi {\left(\boldsymbol{q},\tau \right)}^{\ast}\right]}\cdotp A\left(\boldsymbol{q},\tau \right). $$
(20)

Using Eqs. (8), (19), and (20), it is easy to obtain:

$$ G\left(\boldsymbol{q}-{\boldsymbol{q}}^{\prime },\tau -{\tau}^{\prime}\right)=\frac{\partial^2 lnZ\left({T}_c,J,{J}^{\ast}\right)}{\partial J\left(\boldsymbol{q},\tau \right)\partial J{\left({\boldsymbol{q}}^{\prime },{\tau}^{\prime}\right)}^{\ast }}=\frac{\partial \left\langle \phi \left(\boldsymbol{q},\tau \right)\right\rangle }{\partial J{\left({\boldsymbol{q}}^{\prime },{\tau}^{\prime}\right)}^{\ast }}. $$
(21)

As a quantum extension of the classical Ginzburg number Gi, by using the correlation function (19) we construct an error function of the order parameter ϕ(q, τ) as follows:

$$ {e}^q(D)=\frac{\left|{\int}_0^{\infty } d\tau \int {d}^D\boldsymbol{q}G\left(\boldsymbol{q},\tau \right)\right|}{\int_0^{\infty } d\tau \int {d}^D\boldsymbol{q}\phi {\left(\boldsymbol{q},\tau \right)}^{\ast}\phi \left(\boldsymbol{q},\tau \right)}, $$
(22)

where eq(D) returns to the classical Ginzburg number Gi when ϕ(q, τ) is independent of τ, that is, \( {e}^q(D)=\frac{\left|\int {d}^D\boldsymbol{q}G\left(\boldsymbol{q}\right)\right|}{\int {d}^D\boldsymbol{q}\phi {\left(\boldsymbol{q}\right)}^{\ast}\phi \left(\boldsymbol{q}\right)}={G}_i \) if ϕ(q, τ) = ϕ(q). By Eq. (22), the mean-field approximation is valid if and only if

$$ {e}^q(D)\ll 1. $$
(23)

Therefore, when the inequality (23) breaks down, the nontrivial fixed point (13) holds. To rigidly determine the range of applicability of the nontrivial fixed point (13), we need to explore the physical meaning of the inequality (23). To this end, let us order

$$ M\left({T}_c\right)=\left|{\int}_0^{\infty } d\tau \int {d}^D\boldsymbol{q}G\left(\boldsymbol{q},\tau \right)\right|, $$
(24)
$$ W(t)={\int}_0^{\frac{1}{t}} d\tau \int {d}^D\boldsymbol{q}\phi {\left(\boldsymbol{q},\tau \right)}^{\ast}\phi \left(\boldsymbol{q},\tau \right). $$
(25)

By using Eqs. (24) and (25), Eq. (22) can be written as eq(D) = M(Tc)/W(0). Obviously, we have W(t) ≤ W(0) and M(Tc) ≤ W(0). Since G(q, τ) is the correlation function, M(Tc) actually denotes the magnitude of quantum fluctuations. Thus, the physical meaning of the inequality (23) is that quantum fluctuations can be omitted if and only if their magnitude is extremely small, that is, M(Tc) ≪ W(0). Based on this observation, there should exist a critical magnitude M0 so that when M(Tc) ≥ M0, quantum fluctuations cannot be omitted. This means that the nontrivial fixed point (13) is valid when M(Tc) ≥ M0. To evaluate the value of M(Tc), we introduce an approximation ϕ(q, τ) ≈ 〈ϕ(q, τ)〉 ≈ 〈ϕ(q, τ)〉vac. This approximation has been well-known for evaluating the magnitude of thermal fluctuations when T > 0 [2324].

Proposition 1: If ϕ(q, τ) ≈ 〈ϕ(q, τ)〉 ≈ 〈ϕ(q, τ)〉vac, then the magnitude of quantum fluctuations, M(Tc), yields:

$$ M\left({T}_c\right)={\xi}^2\propto {T}_c^{-2}, $$
(26)

where ξ = (−λ2(Tc))−1/2 denotes the quantum correlation lengthFootnote 2 and 〈ϕ(q, τ)〉vac denotes the vacuum expectation value of 〈ϕ(q, τ)〉.

Proof: see Appendix 2. ■

By Eq. (26), the magnitude M(Tc) and the correlation length ξ grow as Tc decreases, and both of them finally diverge at Tc = 0. This implies that Tc = 0 is a critical point. Since M(Tc) increases as Tc declines, there does exist \( {T}_Q^{\prime } \) so that when \( {T}_c\le {T}_Q^{\prime } \), one has M(Tc) ≥ M0. This means that the nontrivial fixed point (13) is valid when \( {T}_c\le {T}_Q^{\prime } \). To estimate \( {T}_Q^{\prime } \), we construct an index as below:

$$ {E}^q\left(D,t\right)=M\left({T}_c\right)/W(t). $$
(27)

It is easy to check Eq(D, 0) = eq(D) and Eq(D, t) ≥ 0. If we order W(T) = M0, then Eq(D, T) ≥ 1 is equivalent to M(Tc) ≥ M0, where W(t) ≤ W(0) and M(Tc) ≤ W(0) have been used. Thus, the following proposition provides a way for estimating \( {T}_Q^{\prime } \).

Proposition 2: Let us order \( {T}_Q={T}_Q(D)=\mathit{\min}\left\{{T}^{\ast },{T}_Q^{\prime}\right\} \). Eq(D, TQ) ≥ 1 leads to Eq(D, T) ≥ 1.

Proof: Since TQT, we have W(TQ) ≥ W(T), which leads to \( {E}^q\left(D,{T}_Q\right)=\frac{M\left({T}_c\right)}{W\left({T}_Q\right)}\le \frac{M\left({T}_c\right)}{W\left({T}^{\ast}\right)}={E}^q\left(D,{T}^{\ast}\right) \). That is to say, Eq(D, TQ) ≥ 1 leads to Eq(D, T) ≥ 1.■

Since Eq(D, T) ≥ 1 is equivalent to M(Tc) ≥ M0, by the Proposition 2 Eq(D, TQ) ≥ 1 leads to M(Tc) ≥ M0. Therefore, we conclude that the nontrivial fixed point (13) is valid when Eq(D, TQ) ≥ 1. Since TQ is the lower bound of \( {T}_Q^{\prime } \) and the nontrivial fixed point (13) is equivalent to Eq. (16), we have the following criterion:

Criterion A: If Eq(D, TQ) ≥ 1, the parabolic scaling (16) holds for TcTQ.

To estimate TQ by using the Criterion A, we need to calculate Eq(D, TQ). Since M(Tc) has been estimated by Eq. (26), we only calculate the value of W(TQ). As an approximation, we consider that the integral scope of ∫dDqϕ(q, τ)ϕ(q, τ) is up to the correlation length ξ. Thus, by using ϕ(q, τ) ≈ 〈ϕ(q, τ)〉vac, we have:

$$ W\left({T}_Q\right)\approx \frac{1}{T_Q}{\xi}^D{\left|{\left\langle \phi \left(\boldsymbol{q},\tau \right)\right\rangle}_{vac}\right|}^2=\frac{1}{T_Q}{\xi}^D{\rho}_s(0). $$
(28)

Substituting Eqs. (26) and (28) into Eq(D, TQ) yields:

$$ {E}^q\left(D,{T}_Q\right)=\frac{T_Q{\xi}^{2-D}}{\rho_s(0)}. $$
(29)

We now estimate TQ(D) by using Eq. (29). The Criterion A indicates that \( {T}_c=\gamma (D)\cdotp \sqrt{\rho_s(0)} \) holds at Tc = TQ(D), that is, \( {T}_Q(D)=\gamma (D)\cdotp \sqrt{\rho_s(0)} \). Substituting it into Eq(D, TQ) ≥ 1 obtains Eq(D, TQ) = ξ2 − Dγ(D)2/TQ(D) ≥ 1, which indicates:

$$ {T}_Q(D)\le {\xi}^{2-D}\gamma {(D)}^2. $$
(30)

For D = 2, the inequality (30) yields:

$$ {T}_Q(2)\le \gamma {(2)}^2, $$
(31)

which by using the experimental value γ(2) ≈ 4.2 K1/2 yields TQ(2) ≤ 17 K, agreeing with the experimental measure value TQ(2) ≈ 15 K [13].

For D = 3, substituting γ(3) = 0 into the inequality (30) obtains

$$ {T}_Q(3)\le 0, $$
(32)

which indicates that the parabolic scaling (16) holds for TcTQ(3) = 0. That is to say, the mean-field approximation always holds for D = 3. In fact, Tao has pointed out [17] that D = 3 is the upper critical dimension of quantum critical systems and that the mean-field approximation is valid at the upper critical dimension. Therefore, our result for D = 3 agrees with the previous analysis [17].

5 The Two-Class Scaling

By using Abrikosov-Gor’kov’s mean-field theory for superconducting alloys, for dirty BCS superconductors, the relation between Tc and ρs(0) can be derived as [78, 17, 27]:

$$ {T}_c=\alpha \cdotp {\rho}_s(0)+{T}_0. $$
(33)

The derivation for Eq. (33) can be found in Appendix 3. In particular, by using the latest experimental data [28], Khodel et al [27] have produced the correct theoretical value of α. This is an evidence for supporting the linear scaling in Eq. (1) as a result of Abrikosov-Gor’kov’s mean-field theory. By Eq. (1), Eq. (33) holds for TcTM. By the Criterion A, if the mean-field approximation is valid, Eq(D, TQ) ≤ 1 should hold. Using Eq. (27) and TMTQ, it is easy to verify Eq(D, TM) ≤ Eq(D, TQ). This implies that one can estimate TM by using Eq(D, TM) ≤ 1. The following proposition will rigidly confirm this fact.

Proposition 3: Let us order Ω = ∫ dDqϕ(q, τ)ϕ(q, τ). If \( \frac{\mathrm{\partial \Omega }}{\partial \tau }=0 \) and TM > 0, then we have:

$$ {e}^q(D)\ll {E}^q\left(D,{T}_M\right). $$
(34)

Proof: see Appendix 4. ■

Corollary 1: If Eq(D, TM) ≤ 1, then we have eq(D) ≪ 1.

Regarding the Proposition 3, the condition \( \frac{\mathrm{\partial \Omega }}{\partial \tau }=0 \) should approximately hold as long as ϕ(q, τ) ≈ 〈ϕ(q, τ)〉vac is satisfied. Thus, by the Corollary 1, we can replace eq(D) ≪ 1 by Eq(D, TM) ≤ 1 to estimate TM. Since superconducting films imply D = 2, by Eq. (29), we have Eq(2, TM) = TM/ρs(0). By Eq. (1), Tc = α · ρs(0) + T0 holds at Tc = TM. Substituting TM = α · ρs(0) + T0 into Eq(2, TM) ≤ 1 yields \( {E}^q\left(2,{T}_M\right)=\frac{\alpha {T}_M}{T_M-{T}_0}\le 1 \), indicating

$$ {T}_M\ge \frac{T_0}{1-\alpha }, $$
(35)

where we have considered 0 < α < 1 [13] and ρs(0) ≥ 0.

Substituting experimental data α ≈ 0.37 and T0 ≈ 7K into the inequality (35) obtains TM ≥ 11K, which agrees with the experimental value TM ≈ 12K [13].

Using Eqs. (16), (31), (33), and (35), we exactly produce the two-class scaling law for D = 2 as below:

$$ \left\{\begin{array}{cc}{T}_c=\alpha \cdotp {\rho}_s(0)+{T}_0,& {T}_c\ge {T}_M\approx \frac{T_0}{1-\alpha}\\ {}{T}_c=\gamma (2)\cdotp \sqrt{\rho_s(0)},& {T}_c\le {T}_Q\approx \gamma {(2)}^2\end{array}\right., $$
(36)

where \( \gamma (2)=\sqrt{\frac{7\cdotp \zeta (3)\cdotp {\varepsilon}_F}{15\cdotp \pi \cdotp a\cdotp {m}_e}} \).

The theoretical values of γ(2), TQ, and TM have been listed in Table 1. They agree with experimental measure values. In particular, the difference between TM ≈ 11 K and TQ ≈ 17 K implies that the part over [TM, TQ] should be a combination of linear and parabolic scaling. Here we have fitted Eq. (36) to experimental data in the Fig. 1. The accordance between theoretical formula and experimental data is pretty well. Equation (36) is the main result of this paper. It can be rigidly tested by investigating other quasi-two-dimensional BCS-like superconductors.

Table 1 Comparison of theoretical results with experimental measure values [13]
Fig. 1
figure 1

The experimental data from [13] are plotted as black circles, which belong to the Tc interval [5.1 K,  41.6 K]. a The theoretical parabolic scaling (red line) \( {T}_c=4.29\ {K}^{1/2}\bullet \sqrt{\rho_s(0)} \) perfectly fits the experimental data in [5.1 K,  TM], while the linear scaling (blue line) perfectly fits the experimental data in [TQ,  41.6 K], where TM ≈ 11 K and TQ ≈ 17 K, as predicted by Eq. (36). b The theoretical parabolic scaling (red line) \( {T}_c=4.29\ {K}^{1/2}\bullet \sqrt{\rho_s(0)} \) is fitted with the experimental data in the Tc interval [0, 15 K] , where TM ≈ 12 K and TQ ≈ 15 K are experimentally measured [13]

6 Conclusion

In conclusion, by using the BCS theory, we propose a QPF to describe quantum critical behaviors of zero-temperature Cooper pairs. It was recently found that, in the overdoped side of the single-crystal La2 − xSrxCuO4 films, a two-class scaling law emerges as: \( {T}_c=\gamma \cdotp \sqrt{\rho_s(0)} \) for TcTQ and Tc = α · ρs(0) + T0 for TcTM. By using the QPF, we show that the parabolic scaling \( {T}_c=\gamma \cdotp \sqrt{\rho_s(0)} \) can be exactly derived when Tc is sufficiently low, where the theoretical value of γ is exactly calculated as 4.29 K1/2, being in accordance with the experimental measure value γ = (4.2 ± 0.5) K1/2. Furthermore, we show that the linear scaling Tc = α · ρs(0) + T0 is a mean-field behavior of the dirty-limit BCS theory, which lies far beyond the control of the QPF. To determine the range of applicability of the QPF, we extend the classical Ginzburg number to a quantum version. By using the quantum Ginzburg number, we show that the QPF holds for TcTQ, while the mean-field theory holds for TcTM, where theoretical values of TQ and TM are estimated as TQ ≈ 17 K and TM ≈ 11 K, respectively, agreeing with experimental measure values 15 K and 12 K. The high accordance of theoretical values of γ, TQ, and TM with experimental measure results justifies the validity of the QPF. Finally, the QPF predicts that for 2-dimensional overdoped cuprate films, the transition temperature Tc and the quantum correlation length ξ will obey a scaling \( \xi \propto {T}_c^{-\delta } \) with a critical exponent δ being around 1.25. This is a new prediction that can be tested. We propose that one can measure δ by using neutron scattering experiments near Tc = 0, which have been successfully carried out for measuring the critical exponent of the thermal correlation length [26].