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

Number Parity effects in the normal state of SrTiO3SrTiO_{3}

Xing Yang School of Physics and Electronics, Hunan University, Hunan, China, 410082
Physics Department, University of Notre Dame, Notre Dame, Indiana, USA, 45665
   Quanhui Liu School of Physics and Electronics, Hunan University, Hunan, China, 410082    Boldizsár Jankó Physics Department, University of Notre Dame, Notre Dame, Indiana, USA, 45665
Abstract

We study the recently discovered even-odd effects in the normal state of single-electron devices manufactured at strontium titanium oxide/lanthanum aluminum oxide interfaces (STO/LAO). Within the framework of the number parity-projected formalism and a phenomenological fermion-boson model we find that, in sharp contrast to conventional superconductors, the crossover temperature TT^{*} for the onset of number parity effect is considerably larger than the superconducting transition temperature TcT_{c} due to the existence of a pairing gap above TcT_{c}. Furthermore, the finite lifetime of the preformed pairs reduces by several orders of magnitude the effective number of states NeffN_{\rm eff} available for the unpaired quasiparticle in the odd parity state of the Coulomb blockaded STO/LAO island. Our findings are in qualitative agreement with the experimental results reported by Levy and coworkers for STO/LAO based single electron devices.

PACS numbers

73.23.Hk, 73.40.-c, 73.63.-b, 74.25.-q, 74.78.-w

pacs:
73.23.Hk, 73.40.-c, 73.63.-b, 74.25.-q, 74.78.-w
preprint: APS/123-QED

I Introduction

Number parity effects in superconductors were expected as soon as the Bardeen-Cooper-Schrieffer (BCS) microscopic model was developed (Bardeen et al., 1957). Indeed, the BCS ground state corresponds to a coherent superposition of pair states in which the number of particles has even parity and the total number NN is not fixed. Under these circumstances, the charge displacement operator exp(iϕ)exp(i\phi) (canonically conjugate with the number operator N^\hat{N}) has a fixed expectation value, which leads to the common notion that a macroscopic BCS superconductor has a complex order parameter Δ\Delta with a rigid phase ϕ\phi. As soon as the BCS state is projected Schrieffer (1999) onto fixed N, it becomes clear that one has to differentiate between two cases: (a) if the total number N=2nN=2n is even, all particles can participate in pair states and the ground state resembles the usual grand canonical BCS ground state; (b) if N=2n+1N=2n+1 is odd, the ground state will inevitably contain not only pairs, but also an unpaired electron (more precisely: the ground state will contain a Bogoliubov quasiparticle).

Intuitively, one would expect that the NN vs N+1N+1 (even/odd) difference in a superconductor or any kind of paired fermionic state must be experimentally observable only if N is relatively small. Indeed, inspired by the success of the BCS theory, Bohr, Mottelson and Pines Bohr et al. (1958) were the first of many who studied pairing and even-odd effects in nuclear matter, with particle numbers around N102N\sim 10^{2}. It was therefore, even more surprising, when Mooij et al. Geerligs et al. (1990), Tinkham and coworkers Tuominen et al. (1992), as well as Devoret and his colleagues Lafarge et al. (1993a, b) showed experimentally measurable difference between Coulomb blockaded mesoscopic superconducting islands that contain a billion, and a billion plus one electrons. As it turns out, the magnitude of NN was less important. Instead, the quality of the Coulomb blockade turned out to be crucial: the superconducting islands had to be isolated from their environment with ultrasmall tunnel junctions and highly resistive electromagnetic environment, in order to ensure that NN is a fixed, good quantum number.

The pioneering experiments on number parity effects in conventional superconductors were performed on single-electron (SET) devices consisting of lithographically patterned aluminum islands Grabert and Devoret (1992). Even-odd effects emerged below a crossover temperature TT^{*} that was always much lower than the superconducting transition temperature: TTcT^{*}\ll T_{c}. Rather than being directly correlated with TcT_{c}, TT^{*} is set by the experimentally measurable even-odd free energy difference δe/oΔ0kBTlogNeff\delta{\cal F}_{\rm e/o}\sim\Delta_{0}-k_{B}T\log N_{\rm eff}. Here Δ0\Delta_{0} is the low temperature energy gap, and NeffN_{\rm eff} is the effective number of states Tuominen et al. (1992); Lafarge et al. (1993a, b); Jankó et al. (1994) available for the unpaired electron to explore in the odd number parity state of the superconducting island. Within this parity projected framework Tuominen et al. (1992); Jankó et al. (1994) TT^{*} corresponds to the temperature at which δe/o\delta{\cal F}_{\rm e/o} becomes negligibly small: TΔ0/(kBlogNeff)T^{*}\sim\Delta_{0}/(k_{B}\log N_{\rm eff}). For typical device parameters in these early experiments, the crossover temperature was measured to be around T102mKT^{*}\sim 10^{2}{\rm mK} for aluminum island with Tc1KT_{c}\sim 1{\rm K}. Consequently, the effective number of states was typically around Neff104N_{\rm eff}\sim 10^{4}.

The experiments by Levy and his coworkers Cheng et al. (2015) on SET devices constructed on STO/LAO provided experimental evidence for a spectacular departure from the conventional number parity effects described above. Levy and his colleagues detected T900mKT^{*}\sim 900{\rm mK}, much higher than the superconducting transition temperature Tc300mKT_{c}\sim 300{\rm mK} measured for these devices. Even-odd effects remained detectable well into the ”normal” phase of the superconductor, and persisted in magnetic fields B14TB^{*}\sim 1\sim 4{\rm T}, much higher than the upper critical field of the device. Furthermore, the extracted NeffN_{\rm eff} is also drastically different: Neff23N_{\rm eff}\sim 2-3.

A possible and relatively straightforward interpretation of novel experimental developments suggest that preformed pairs Geshkenbein et al. (1997); Seo et al. (2019); Tagliavini et al. (2016) persist into the normal state of STO-LAO well above the superconducting transition temperature. Consequently, fundamental changes must be made to the theoretical description of the number parity effects in this novel preformed pair phase. This paper is devoted to the presentation of a phenomenological theoretical framework aimed at providing a description of number parity effects in the normal phase of STO/LAO devices. Given the fact that the details of the microscopic mechanism behind the superconducting and preformed pair state of STO/LAO are not yet established, we use a phenomenological fermion-boson model Alexandrov (1996) that allows us to describe a normal phase where both pairs and unpaired particles are present. Furthermore, the model allows pairs to decay into unpaired particles, and particles to form pairs. This theoretical picture provides in a natural way a finite pair lifetime Cheng et al. (2015) in the preformed pair state. We find, after performing the number parity projection developed earlier by Ambegaokar, Smith and one of us Jankó et al. (1994), that the finite pair lifetime has drastic effect on the magnitude of NeffN_{\rm eff}. In fact, as we will show in detail below, the theoretical framework we develop in this paper can reproduce not only TTcT^{*}\gg T_{c}, but also Neff𝒪(1)N_{\rm eff}\sim\mathcal{O}(1).

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 1: Qualitative sketch of the momentum space and density of states for (a) the density of states in the metallic normal state; (b) a conventional superconductor with a superconducting gap, and (c) an unconventional superconductor with a zero-width gap at Fermi level. For the cases (b) and (c), even-odd effect can be expected because of the presence of the superconducting gaps.

Generally speaking, the materials for making the single-electron devices can be separated into three categories (see Fig. 1): gapless materials, BCS and unconventional superconductors. Their density of states are shown, respectively, in panels (a), (b) and (c) of Fig. 1. While the single electron transistors with ultrasmall islands and discrete energy spectrum have also been investigated extensively Smith and Ambegaokar (1996); von Delft et al. (1996), we will not discuss this regime here. According to our calculations, the density of states at the Fermi level should be vanishingly small in order to obtain a finite even/odd free energy difference. As a result, the single electron transistors made from BCS superconductors and unconventional superconductors (as shown in panels (b) and (c)) are expected to show even-odd effects, and a superconducting gap above TcT_{c} is necessary to cause TTcT^{*}\gg T_{c}. The effective excitation number for the unpaired electrons in the odd parity states is highly dependent on the density of states at E=ΔE=\Delta, since the smallest excitation energy is assumed to be Δ\Delta Tuominen et al. (1992); Lafarge et al. (1993a, b). In BCS superconductors (see panel (b)), the density of states at E=ΔE=\Delta is known to have a van Hove singularity, and this results in a large Neff104N_{\rm eff}\sim 10^{4}. In unconventional superconductors (see panel (c)), the van Hove singularity is broadened by the presence of low energy quasiparticles, which results in a small Neff𝒪(1)N_{\rm eff}\sim\mathcal{O}(1).

Several possible microscopic superconducting mechanisms of the electron system at the STO/LAO interface have been proposed recently by different groups. Ruhman and Lee Ruhman and Lee (2016) suggested on the plasmon-induced superconducting mechanism. A nonperturbative approach within the plasmon model is being developed by Edelman and Littlewood Edelman and Littlewood . Kedem et al. Edge et al. (2015); Kedem (2018) related the mechanism to the ferroelectric mode. Arce-Gamboa and Guzmán-Verri Arce-Gamboa and Guzmán-Verri (2018) discovered the influence of strain on the ferroelectric mode and obtained the dependence of superconducting transition temperature on cation doping. On the contrary, Wölfle and Balastsky Wölfle and Balatsky (2018) proposed the transverse optical phonons may be the glue for electron pairing. While these theories can explain the origin of superconducting gap selfconsistently, some important experimental facts in remain unexplained. First of all, the pairing gap should persist above TcT_{c}. This is very important for the even-odd effects above TcT_{c} in the single electron transistorsRichter (2013). Next, the van Hove singularity in the density of states should be broadened out. This will lead to a small NeffN_{\rm eff} consistent with the experiments of Levy et al Cheng et al. (2015). The detailed discussion of these results will be presented in the sections below.

This paper is organized as follows. In Section II, two important physical quantities, the even/odd free energy difference and the effective excitation number for the unpaired electron in the odd parity state, are related to the density of states within the phenomenological Dynes formula. In Section III, the boson-fermion model is introduced, and the analytic form of its electron Green’s function is provided. The density of states and the physical quantities of the even-odd effect predicted by the model are calculated in Section IV. Finally, we present our conclusions in Section V.

II even/odd free energy difference and effective excitation number for the unpaired electron

II.1 Even/odd free energy difference

The electron system of the quantum dot at the STO/LAO interface is described by a general Hamiltonian H^\hat{H}. Within the number parity projection formalism Jankó et al. (1994), the canonical partition function with even/odd number parity is :

Ze/o=Tr{1±(1)N^2eβ(H^μN^)}Z_{e/o}={\rm Tr}\bigg{\{}{\frac{1\pm(-1)^{\hat{N}}}{2}}e^{-\beta(\hat{H}-\mu\hat{N})}\bigg{\}} (1)

where the symbol e/oe/o corresponds to the even/odd parity, N^\hat{N} is the electron number operator, while β=1kBT\beta=\frac{1}{k_{B}T} where kBk_{B} is Boltzmann constant and TT is the temperature.

From Eq. 1, the difference between the free energy of a system with an odd and even number of particles is

FoFe=1βln[1+(1)N^1(1)N^],F_{o}-F_{e}=\frac{1}{\beta}\ln\bigg{[}\frac{1+\langle(-1)^{\hat{N}}\rangle}{1-\langle(-1)^{\hat{N}}\rangle}\bigg{]}, (2)

where Tr{eβ(H^μN^)}/Z\langle...\rangle\equiv{\rm Tr}\{e^{\beta(\hat{H}-\mu\hat{N})}...\}/Z and Z=Tr{eβ(H^μN^)}Z={\rm Tr}\{e^{-\beta(\hat{H}-\mu\hat{N})}\}. The expectation value (1)N^\langle(-1)^{\hat{N}}\rangle is the parameter that signals the presence or absence of even-odd effects. When (1)N^=0\langle(-1)^{\hat{N}}\rangle=0, even-odd effects will not be observable. Let us assume that the Hamiltonian can be expressed in a more compact form H^=𝐤,σe𝐤c^𝐤,σc^𝐤,σ\hat{H}=\sum_{\mathbf{k},\sigma}{e_{\mathbf{k}}\hat{c}^{\dagger}_{\mathbf{k},\sigma}\hat{c}_{\mathbf{k},\sigma}}. From the above, (1)N^\langle(-1)^{\hat{N}}\rangle is

(1)N^=𝐤(1eβ(e𝐤μ))2(1+eβ(e𝐤μ))2=𝐤tanh2(β(e𝐤μ)2).\langle(-1)^{\hat{N}}\rangle=\prod_{\mathbf{k}}\frac{(1-e^{\beta(e_{\mathbf{k}}-\mu)})^{2}}{(1+e^{\beta(e_{\mathbf{k}}-\mu)})^{2}}=\prod_{\mathbf{k}}\tanh^{2}(\frac{\beta(e_{\mathbf{k}}-\mu)}{2}). (3)

If AA is defined as eA(1)N^e^{A}\equiv\langle(-1)^{\hat{N}}\rangle, then

A=2𝐤ln|tanhβe𝐤2|=2+D(E)ln|tanhβE2|dEA=2\sum_{\mathbf{k}}\ln\bigg{|}\tanh\frac{\beta e_{\mathbf{k}}}{2}\bigg{|}=2\int^{+\infty}_{-\infty}D(E)\ln\bigg{|}\tanh\frac{\beta E}{2}\bigg{|}dE\\ (4)

where D(E)D(E) is the density of states. Notice that the factor ln|tanhβE2|\ln|\tanh\frac{\beta E}{2}| in the integrand is divergent when E=0E=0. This suggests that an energy gap is necessary for a system to show even-odd effects. In the absence of a pairing gap AA is a large negative number and eA0e^{A}\approx 0. At the interface of STO/LAO, the even-odd effects appear above the superconducting transition temperature TcT_{c}. This implies the existence of a pairing gap above TcT_{c}. Scanning tunneling spectroscopy experiments also show that an energy gap persists above TcT_{c}Richter (2013). It is one of the requirements for a system showing number-parity effects. Moreover, the factor ln|tanhβE2|\ln|\tanh\frac{\beta E}{2}| turns to be zero when EEFE\gg E_{F}, this suggests the part with the high energy does not contribute to the integral AA. On the other hand, the density of states near Fermi level (gap states) can increase the value of |A||A| greatly, and the even-odd effect parameter (1)N^=eA=e|A|1\langle(-1)^{\hat{N}}\rangle=e^{A}=e^{-|A|}\ll 1 is reduced accordingly. In this sense, the emergence of the gap states can weaken even-odd effects.

II.2 Effective excitation number for the unpaired electron

With the assumption that the smallest excitation energy for electrons is Δ\Delta, we can calculate the effective excitation number for the unpaired electron with the formula Tuominen et al. (1992); Lafarge et al. (1993a, b):

Neff=Δ0D(E)exp(β(EΔ))𝑑E.N_{\rm eff}=\int^{\infty}_{\Delta_{0}}D(E)\exp(-\beta(E-\Delta))dE. (5)

In Eq. 5, the density of states at E=ΔE=\Delta contributes most to the effective excitation number for the unpaired electron in the odd parity states. If it is assumed that the van Hove singularity in D(E)D(E) exists, it can easily produce a large Neff104N_{\rm eff}\sim 10^{4} or more in BCS superconductors. However, the experiments Cheng et al. (2015) at the interface of STO/LAO discover a very small Neff1N_{\rm eff}\sim 1. This indicates that the van Hove singularity was broadened out in the density of states.

II.3 Even-odd effect phenomenology with the Dynes formula

As we can see in the above discussion, the even-odd effects are related to the density of states of the small island in the single electron transistors. In order to reproduce the experimental results, the van Hove singularity should, at least, be broadened. This can be provided by the lifetime effects of electron pairs. Here we adopt the phenomenological Dynes formula and calculate the even/odd free energy difference δFe/o\delta F_{e/o} and the effective excitation number for the unpaired electron NeffN_{\rm eff}

To be explicit, we will use the following form for the density of statesDynes et al. (1978)

Dd(E)=Dn(0)Re|EiΓ|(EiΓ)2Δ2D_{d}(E)=D_{n}(0)\mathrm{Re}\frac{|E-i\Gamma|}{\sqrt{(E-i\Gamma)^{2}-\Delta^{2}}} (6)

where Dn(0)D_{n}(0) is the density of states in the normal state, Γ\Gamma is the phenomenological imaginary part of energy, and Δ\Delta is the superconducting energy gap. With Eqs. 2-5, the even/odd free energy difference δFe/o\delta F_{e/o} and the effective excitation state number for the unpaired electron NeffN_{\rm eff} are calculated and plotted in Fig. 2. As we can see, NeffN_{\rm eff} reduces to 1\sim 1, and the even/odd free energy difference is finite provided that the superconducting gap Δ\Delta does not close. This result is independent of any microscopic model. In order for the pairing-induced even-odd effect to be visible, the density of states at the Fermi level must vanish, and broadening Γ\Gamma has to be small compared to the gap Δ0\Delta_{0}

Refer to caption
Figure 2: Γ\Gamma versus the even/odd free energy difference δFe/o\delta F_{e/o} and the number of the effective excitation states for the unpaired quasiparticles, NeffN_{\rm eff}.

III The boson-fermion Model

As mentioned in the introduction, there is no consensus yet on the microscopic theory of superconductivity in STO. In order to reproduce a Dynes-like density of states, we turn to the phenomenological boson-fermion model. For a single band model, electrons are assumed to have a Bogoliubov quasiparticle dispersion E𝐤=ϵ𝐤2+Δ2E_{\mathbf{k}}=\sqrt{\epsilon^{2}_{\mathbf{k}}+\Delta^{2}}, where ϵ𝐤=2k22mμF\epsilon_{\mathbf{k}}=\frac{\hbar^{2}k^{2}}{2m^{*}}-\mu_{F}. The Hamiltonian of the electrons can be written as:

H^0e=𝐤,σE𝐤c^𝐤,σc^𝐤,σ.\hat{H}_{0e}=\sum_{\mathbf{k},\sigma}{E_{\mathbf{k}}\hat{c}^{\dagger}_{\mathbf{k},\sigma}\hat{c}_{\mathbf{k},\sigma}}. (7)

The superconducting gap of the 2D electron system at the interface of STO/LAO, Δ\Delta vanishes at Ts300mKT_{s}\sim 300{\rm mK}, and it turns into superconducting state at Tc190mKT_{c}\sim 190{\rm mK} Richter (2013). This suggests between TcT_{c} and TsT_{s}, the superconducting phase is destroyed, but the superconducting gap is preserved. This regime corresponds to the preformed pair state. The present model is devoted to studying the preformed pair state and superconducting state. Notice that the coherence length of pairs is 70100nm\sim{\rm 70-100nm} in (001)-STO/LAO and 4075nm{\rm 40-75nm} for (011)-STO/LAO Herranz et al. (2015), which is very small compared to the coherence length in conventional superconductors. In order to give an approximate description of preformed pairs, we introduce a bosonic field b^𝐪\hat{b}_{\mathbf{q}} with elementary charge unit 2e-2e. For a small momentum 𝐪\mathbf{q}, the dispersion of the pairs is approximated as ξ𝐪=ξ0+v|q|μb\xi_{\mathbf{q}}=\xi_{0}+\hbar v|q|-\mu_{b} Schrieffer (1999), and the Hamiltonian for the bare bosonic field is:

H^0p=𝐪ξ𝐪b^𝐪b^𝐪\hat{H}_{0p}=\sum_{\mathbf{q}}{\xi_{\mathbf{q}}\hat{b}^{\dagger}_{\mathbf{q}}\hat{b}_{\mathbf{q}}} (8)

where b^𝐪\hat{b}^{\dagger}_{\mathbf{q}} and b^𝐪\hat{b}_{\mathbf{q}} are defined to commute with c^𝐤,σ\hat{c}^{\dagger}_{\mathbf{k},\sigma} and c^𝐤,σ\hat{c}_{\mathbf{k},\sigma}. The interaction Hamiltonian between the fermions and bosons is assumed to be:

H^1=𝐤,𝐪V1(𝐪)n0b^𝐪c^𝐤+𝐪2c^𝐤+𝐪2+H.c.\hat{H}_{1}=\sum_{\mathbf{k},\mathbf{q}}{\frac{V_{1}(\mathbf{q})}{\sqrt{n_{0}}}\hat{b}^{\dagger}_{\mathbf{q}}\hat{c}_{-\mathbf{k}+\frac{\mathbf{q}}{2}\downarrow}\hat{c}_{\mathbf{k}+\frac{\mathbf{q}}{2}\uparrow}+H.c.} (9)

The total Hamiltonian is

H^=H^0+H^1\hat{H}=\hat{H}_{0}+\hat{H}_{1} (10)

where

H^0=H^0e+H^0p.\hat{H}_{0}=\hat{H}_{0e}+\hat{H}_{0p}. (11)

The total particle number is defined as N^=𝐤,σc^𝐤,σc^𝐤,σ+2𝐪b^𝐪b^𝐪\hat{N}=\sum_{\mathbf{k},\sigma}{\hat{c}^{\dagger}_{\mathbf{k},\sigma}\hat{c}_{\mathbf{k},\sigma}}+2\sum_{\mathbf{q}}{\hat{b}^{\dagger}_{\mathbf{q}}\hat{b}_{\mathbf{q}}}, and it can be proven that [N^,H^]=0[\hat{N},\hat{H}]=0. The first order approximation of the self-energy is

Σ(𝐤,ω)=12L|V1(𝐪)|2d𝐪2πn01ω(ξ𝐪E𝐪𝐤)+iη×(1eβξ𝐪1+1eβE𝐪𝐤+1)\begin{split}&\Sigma(\mathbf{k},\omega)=\frac{1}{\hbar^{2}}\int\frac{L|V_{1}(\mathbf{q})|^{2}d\mathbf{q}}{2\pi n_{0}}\frac{1}{\omega-\frac{(\xi_{\mathbf{q}}-E_{\mathbf{q}-\mathbf{k}})}{\hbar}+i\eta}\\ &\times\bigg{(}\frac{1}{e^{\beta\xi_{\mathbf{q}}}-1}+\frac{1}{e^{\beta E_{\mathbf{q}-\mathbf{k}}}+1}\bigg{)}\\ \end{split} (12)

where LL is the length of the quantum dot in the middle of the single electron transistor. Notice that the superconductivity of the STO/LAO system is considered to be one-dimensional Cheng et al. (2015), which makes our proposed theory to be one-dimensional as well. Let us now introduce a momentum dependent interaction kernel V1(q)V_{1}(q) as an example interaction that reproduces a Dynes-like density of states:

V1=Vc(ξ𝐪E𝐪𝐤)2(ξ𝐪E𝐪𝐤)2+Δ02V_{1}=V_{c}\sqrt{\frac{(\xi_{\mathbf{q}}-E_{\mathbf{q}-\mathbf{k}})^{2}}{(\xi_{\mathbf{q}}-E_{\mathbf{q}-\mathbf{k}})^{2}+\Delta_{0}^{2}}} (13)

where VcV_{c} is the strength of the coupling, and n0n_{0} is the total number of quasiparticles in the quantum dot. n0n_{0} is around 500500 Cheng et al. (2015). Notice that the factor ξ𝐪E𝐪𝐤\xi_{\mathbf{q}}-E_{\mathbf{q}-\mathbf{k}} in fact is equivalent to the frequency of the electron in Green’s function ω\omega. The calculation of the self-energy is presented in Appendix A.

IV even-odd effect within the boson-fermion model

With Eqs. 17,18, the one-particle Green’s function can be written as:

G(𝐤,ω)=1ωϵ𝐤ΣiΓG(\mathbf{k},\omega)=\frac{1}{\omega-\epsilon_{\mathbf{k}}-\Sigma^{\prime}-i\Gamma} (14)

where ϵ𝐤=Δ2+(2(k2kF2)2m)2Δ2+(vF(kkF))2\epsilon_{\mathbf{k}}=\sqrt{\Delta^{2}+(\frac{\hbar^{2}(k^{2}-k_{F}^{2})}{2m^{*}})^{2}}\approx\sqrt{\Delta^{2}+(\hbar v^{*}_{F}(k-k_{F}))^{2}}, vFkFmv^{*}_{F}\equiv\frac{\hbar k_{F}}{m^{*}}, Σ=Σ𝐤(ω)|k=kF\Sigma^{\prime}=\Sigma^{\prime}_{\mathbf{k}}(\omega)|_{k=k_{F}}, Γ=Γ𝐤(ω)|k=kF\Gamma=\Gamma_{\mathbf{k}}(\omega)|_{k=k_{F}}. Numerical calculations show that ΣΔ\Sigma^{\prime}\ll\Delta at low temperature and consequently Σ\Sigma^{\prime}is negligible.

The density of states is

D(ω)=LA(𝐤,w)𝑑𝐤D(\omega)=\int_{-\infty}^{\infty}LA(\mathbf{k},w)d\mathbf{k} (15)

where A(𝐤,ω)=1πIm(G(𝐤,ω))A(\mathbf{k},\omega)=-\frac{1}{\pi}{\rm Im}(G(\mathbf{k},\omega)) and L500nmL\approx 500{\rm nm} is the length of the small island in the middle of the single electron transistor. Notice that only ϵ𝐤\epsilon_{\mathbf{k}} is dependent on the momentum 𝐤\mathbf{k} in the spectral weight function A(𝐤,ω)A(\mathbf{k},\omega). This allows us to deduce an exact result in the mathematical expression of D(ω)D(\omega) with the residue theorem (see Appendix B).

As shown in Fig. 3, the decay and formation of the electron pairs produces many gap states and broadens the van Hove singularity in the density of states. The effects can reduce the even/odd free energy difference δe/o\delta{\cal F}_{\rm e/o}, and the effective excitation number for the unpaired electron in the odd parity state NeffN_{\rm eff}, and this can be measured in experiments. In addition, zero superconducting gap makes a finite spectral function at Fermi level, and in that case, the density of states, D(ω)D(\omega), is finite at the Fermi level. This can destroy the even-odd effects. As shown in Fig. 3, the results on the even-odd effects calculated by the boson-fermion model is very similar to those we obtained from Dynes’ model density of states.

Refer to caption
Figure 3: The density of states, D(ω)D(\omega), is plotted along with the variation of VcV_{c}. D(ω)D(\omega) and VcV_{c} are plotted in atomic units.
Refer to caption
Figure 4: The even-odd free energy difference and the effective excitation number for the unpaired electron versus VcV_{c}. It is plotted in atomic units. m=mem^{*}=m_{e}, n0=500n_{0}=500, L=530nmL=530{\rm nm}, vF=8.8×103m/sv^{*}_{F}=8.8\times 10^{3}{\rm m/s} and v=0.073cv=0.073c,where cc is the speed of light in vacuum.

V Conclusion

In the present paper, we argue that the even-odd effects seen in the normal state of STO originate from superconducting preformed pairing. This in turn imposes severe constraints on the density of states and consequently any microscopic model aimed at explaining the superconducting and normal state of STO/LAO. First, the density of states at Fermi level should be zero below and above TcT_{c}. A superconducting gap is required for the electron system to demonstrate even-odd effects above TcT_{c}. Next, the gap states are necessary to reduce the even/odd free energy difference and weaken the even-odd effects. Finally, the van Hove singularity needs to be broadened out in order to obtain a small NeffN_{\rm eff}. These constraints for the density of states are not immediately satisfied by most current microscopic theories. On the other hand, the theories are successful in explaining the microscopic origin of the superconducting gap Δ\Delta and the dispersion of Bogoliubov quasiparticles.

The broadening of the van Hove singularity in the density of states may be a fingerprint of the lifetime effects of electron pairs. The signals in the single electron transistor experiments are very sensitive in detecting the lifetime effects, as well as the existence of the superconducting gap, the gap states and the broadening of the van Hove singularity. Moreover, compared to the experimental condition of scanning tunneling spectroscopy, that of the single electron devices can be relatively more easily satisfied in some strongly interacting electron systems. Furthermore, the decay and formation of electron pairs may widely exist in many different types of superconductors, including BCS superconductors. The application of the single electron transistor devices to study novel superconductors is therefore very promising.

The microscopic origin of phenomenological interaction potential is still unknown. Phonon-electron interactions, electron-electron interactions, etc. may participate in the actual microscopic mechanismYang and Liu . Further theoretical and experimental investigations are needed to elucidate the detailed microscopic model of the superconducting and normal state of SrTiO3.

Acknowledgements.
One of us (X. Y.) gratefully acknowledges the support from China Scholarship Council under Grant No. 201506130054. We thank Peter B. Littlewood, Alexander Edelman, Anthony Ruth and Xiaoyu Ma for helpful discussions. This work is financially supported by National Natural Science Foundation of China under Grant No. 11675051.

Appendix A Calculations of the self-energy

The decay rate of quasiparticles is defined to be Γ=Im(Σ(𝐤,ω))\Gamma=\mathrm{Im}(\Sigma(\mathbf{k},\omega)). In order to facilitate the calculations, it is assumed that Σ(𝐤,ω)=Σ(𝐤,ω)||𝐤|=kF\Sigma(\mathbf{k},\omega)=\Sigma(\mathbf{k},\omega)\big{|}_{|\mathbf{k}|=k_{F}}. From Eq. 18 and the above approximations, the decay rate of electron pairs is

Γ(ω)=12L|V1(𝐪)|2d𝐪2n0δ(ω(ξ𝐪E𝐪))×(1eβξ𝐪1+1eβE𝐪+1).\begin{split}&\Gamma(\omega)=\frac{1}{\hbar^{2}}\int\frac{L|V_{1}(\mathbf{q})|^{2}d\mathbf{q}}{2n_{0}}\delta(\omega-\frac{(\xi_{\mathbf{q}}-E_{\mathbf{q})}}{\hbar})\\ &\times(\frac{1}{e^{\beta\xi_{\mathbf{q}}}-1}+\frac{1}{e^{\beta E_{\mathbf{q}}}+1}).\\ \end{split} (16)

When qq is very small and vFqΔ\hbar v^{*}_{F}q\ll\Delta, the equation ω(ξ𝐪E𝐪)=0\omega-\frac{(\xi_{\mathbf{q}}-E_{\mathbf{q})}}{\hbar}=0 becomes ω(ξ0+v|q|μb)+Δ0\omega-(\xi_{0}+v|q|-\mu_{b})+\Delta\approx 0 where \hbar is assumed to be unity. The chemical potential of bosons is set to be ξ0μbΔ\xi_{0}-\mu_{b}\approx\Delta. This leads to the solution that ωv|q|\omega\approx v|q|. The frequency of the electrons ω\omega is in the order of Δ\sim\Delta. This self-consistently proves that qq is very small and vFqΔ\hbar v^{*}_{F}q\ll\Delta. With the results, we obtain

Γ(ω)L|V1(ωv)|2vn0(1eβ(Δ+ω)1+1eβΔ+1).\Gamma(\omega)\approx\frac{L|V_{1}(\frac{\omega}{v})|^{2}}{vn_{0}}(\frac{1}{e^{\beta(\Delta+\omega)}-1}+\frac{1}{e^{\beta\Delta}+1}). (17)

With the particle-hole parity symmetry and Kramers-Krönig relation, the real part of the self-energy is

Σ(ω)=Re(Σ(k=kF,ω))=2ωπP0Γ(ω)ω2ω2𝑑ω\Sigma^{\prime}(\omega)=\mathrm{Re}(\Sigma(k=k_{F},\omega))=\frac{2\omega}{\pi}P\int_{0}^{\infty}\frac{\Gamma(\omega^{\prime})}{\omega^{\prime 2}-\omega^{2}}d\omega^{\prime} (18)

Numerical calculations show that the real part of the self-energy has a negligible effect in generating the predicted density of states.

Appendix B Calculations of the Density of States

The denominator of the spectral weight function is (wΔ2+(vF(kkF))2Σ)2+Γ2(w-\sqrt{\Delta^{2}+(\hbar v^{*}_{F}(k-k_{F}))^{2}}-\Sigma^{\prime})^{2}+\Gamma^{2}. If the denominator equals to zero, there are four solutions of the momentum kk that two solutions are in upper half-plane of the complex plane of kk and two solutions are in the lower half. Furthermore, there are four different cases, if we set a=(ωΣ)2Γ2Δ2,b=2Γ(ωΣ)a=(\omega-\Sigma^{\prime})^{2}-\Gamma^{2}-\Delta^{2},b=2\Gamma(\omega-\Sigma^{\prime})

{a>0,b>0.\romannum1a>0,b<0.\romannum2a<0,b>0.\romannum3a<0,b<0.\romannum4.\left\{\begin{aligned} a>0,b>0...................\romannum{1}~{}~{}\\ a>0,b<0...................\romannum{2}~{}\\ a<0,b>0...................\romannum{3}\\ a<0,b<0...................\romannum{4}.\end{aligned}\right.

For case \romannum1, four solutions of the momentum kk in the upper half-plane are

k1\displaystyle k_{1} =kF1vFReθi2\displaystyle=k_{F}-\frac{1}{v^{*}_{F}}\sqrt{R}e^{\frac{-\theta i}{2}} (19)
k2\displaystyle k_{2} =kF+1vFReθi2\displaystyle=k_{F}+\frac{1}{v^{*}_{F}}\sqrt{R}e^{\frac{\theta i}{2}} (20)

where θ=Arctanba,R=a2+b2\theta=Arctan\frac{b}{a},R=\sqrt{a^{2}+b^{2}}. For case \romannum2,

k1=kF+1vFReθi2\displaystyle k_{1}=k_{F}+\frac{1}{v^{*}_{F}}\sqrt{R}e^{\frac{-\theta i}{2}} (21)
k2=kF1vFReθi2.\displaystyle k_{2}=k_{F}-\frac{1}{v^{*}_{F}}\sqrt{R}e^{\frac{\theta i}{2}}. (22)

For case \romannum3,

k1=kF1vFRe(θ+π)i2\displaystyle k_{1}=k_{F}-\frac{1}{v^{*}_{F}}\sqrt{R}e^{\frac{-(\theta+\pi)i}{2}} (23)
k2=kF+1vFRe(θ+π)i2.\displaystyle k_{2}=k_{F}+\frac{1}{v^{*}_{F}}\sqrt{R}e^{\frac{(\theta+\pi)i}{2}}. (24)

For case \romannum4,

k1=kF1vFRe(θπ)i2\displaystyle k_{1}=k_{F}-\frac{1}{v^{*}_{F}}\sqrt{R}e^{\frac{(\theta-\pi)i}{2}} (25)
k2=kF+1vFRe(θ+π)i2.\displaystyle k_{2}=k_{F}+\frac{1}{v^{*}_{F}}\sqrt{R}e^{\frac{(-\theta+\pi)i}{2}}. (26)

After applying Jordan’s lemma and the residue theorem, we obtain the density of states, for case \romannum1,

D(ω)=2LvF[ωΣRcos(θ2)+ΓRsin(θ2)].D(\omega)=\frac{2L}{v^{*}_{F}}\bigg{[}\frac{\omega-\Sigma^{\prime}}{\sqrt{R}}\cos\bigg{(}\frac{\theta}{2}\bigg{)}+\frac{\Gamma}{\sqrt{R}}\sin\bigg{(}\frac{\theta}{2}\bigg{)}\bigg{]}. (27)

For case \romannum2,

D(ω)=2LvF[ωΣRcos(θ2)+ΓRsin(θ2)].D(\omega)=-\frac{2L}{v^{*}_{F}}\bigg{[}\frac{\omega-\Sigma^{\prime}}{\sqrt{R}}\cos\bigg{(}\frac{\theta}{2}\bigg{)}+\frac{\Gamma}{\sqrt{R}}\sin\bigg{(}\frac{\theta}{2}\bigg{)}\bigg{]}. (28)

For case \romannum3,

D(ω)=2LvF[ωΣRcos(θ+π2)+ΓRsin(θ+π2)].D(\omega)=\frac{2L}{v^{*}_{F}}\bigg{[}\frac{\omega-\Sigma^{\prime}}{\sqrt{R}}\cos\bigg{(}\frac{\theta+\pi}{2}\bigg{)}+\frac{\Gamma}{\sqrt{R}}\sin\bigg{(}\frac{\theta+\pi}{2}\bigg{)}\bigg{]}. (29)

For case \romannum4,

D(ω)=2LvF[ωΣRcos(θπ2)+ΓRsin(θπ2)].D(\omega)=-\frac{2L}{v^{*}_{F}}\bigg{[}\frac{\omega-\Sigma^{\prime}}{\sqrt{R}}\cos\bigg{(}\frac{\theta-\pi}{2}\bigg{)}+\frac{\Gamma}{\sqrt{R}}\sin\bigg{(}\frac{\theta-\pi}{2}\bigg{)}\bigg{]}. (30)

References

  • Bardeen et al. (1957) J. Bardeen, L. Cooper,  and J. Schrieffer, Phys. Rev. 108, 1175 (1957).
  • Schrieffer (1999) J. Schrieffer, Theory of Superconductivity (Benjamin-Cummings Publishing Company, 1999).
  • Bohr et al. (1958) A. Bohr, B. R. Mottelson,  and D. Pines, Phys. Rev. 110, 936 (1958).
  • Geerligs et al. (1990) L. J. Geerligs, V. F. Anderegg, J. Romijn,  and J. E. Mooij, Phys. Rev. Lett. 65, 377 (1990).
  • Tuominen et al. (1992) M. T. Tuominen, J. M. Hergenrother, T. S. Tighs,  and M. Tinkham, Phys. Rev. Lett. 69, 1997 (1992).
  • Lafarge et al. (1993a) P. Lafarge, P. Joyez, D. Esteve, C. Urbina,  and M. H. Devoret, Phys. Rev. Lett. 70, 994 (1993a).
  • Lafarge et al. (1993b) P. Lafarge, P. Joyez, D. Esteve, C. Urbina,  and M. H. Devoret, Nature 365, 422 (1993b).
  • Grabert and Devoret (1992) H. Grabert and M. H. Devoret, Single Charge Tunneling: Coulomb Blockade Phenomena In Nanostructures (Springer, 1992).
  • Jankó et al. (1994) B. Jankó, A. Smith,  and V. Ambegaokar, Phys. Rev. B 50, 1152 (1994).
  • Cheng et al. (2015) G. Cheng, M. Tomczyk, S. Lu, J. P. Veazey, M. Huang, P. Irvin, S. Ryu, H. Lee, C. B. Eom, C. S. Hellberg,  and J. Levy, Nature 521, 196 (2015).
  • Geshkenbein et al. (1997) V. B. Geshkenbein, L. B. Ioffe,  and A. I. Larkin, Physical Review B 55, 3173 (1997).
  • Seo et al. (2019) Y. I. Seo, W. J. Choi, S.-i. Kimura,  and Y. S. Kwon, Scientific Reports 9, 1 (2019).
  • Tagliavini et al. (2016) A. Tagliavini, M. Capone,  and A. Toschi, Phys. Rev. B 94, 155114 (2016).
  • Alexandrov (1996) A. S. Alexandrov, Journal of Physics: Condensed Matter 8, 6923 (1996).
  • Smith and Ambegaokar (1996) R. A. Smith and V. Ambegaokar, Phys. Rev. Lett. 77, 4962 (1996).
  • von Delft et al. (1996) J. von Delft, A. D. Zaikin, D. S. Golubev,  and W. Tichy, Phys. Rev. Lett. 77, 3189 (1996).
  • Ruhman and Lee (2016) J. Ruhman and P. A. Lee, Phys. Rev. B 94, 224515 (2016).
  • (18) A. Edelman and P. B. Littlewood, To be published .
  • Edge et al. (2015) J. M. Edge, Y. Kedem, U. Aschauer, N. A. Spaldin,  and A. V. Balatsky, Phys. Rev. Lett. 115, 247002 (2015).
  • Kedem (2018) Y. Kedem, Phys. Rev. B 98, 220505 (2018).
  • Arce-Gamboa and Guzmán-Verri (2018) J. R. Arce-Gamboa and G. G. Guzmán-Verri, Phys. Rev. Materials 2, 104804 (2018).
  • Wölfle and Balatsky (2018) P. Wölfle and A. V. Balatsky, Phys. Rev. B 98, 104505 (2018).
  • Richter (2013) C. Richter, Experimental Investigation of Electronic and Magnetic Properties of LaAlO3-SrTiO3 Interfaces, doctoralthesis, Universität Augsburg, Augsburg (2013).
  • Dynes et al. (1978) R. C. Dynes, V. Narayanamurti,  and J. P. Garno, Phys. Rev. Lett. 41, 1509 (1978).
  • Herranz et al. (2015) G. Herranz, G. Singh, N. Bergeal, A. Jouan, J. Lesueur, J. Gázquez, M. Varela, M. Scigaj, N. Dix, F. Sánchez,  and J. Fontcuberta, Nature Communications 6, 6028 (2015).
  • (26) X. Yang and Q. Liu, To be published .